30.06.2013 Views

Wind Hazard Risk Assessment and Management for Structures

Wind Hazard Risk Assessment and Management for Structures

Wind Hazard Risk Assessment and Management for Structures

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

<strong>Wind</strong> <strong>Hazard</strong> <strong>Risk</strong> <strong>Assessment</strong> <strong>and</strong> <strong>Management</strong><br />

<strong>for</strong> <strong>Structures</strong><br />

Siu Chung Yau<br />

A Dissertation<br />

Presented to the Faculty<br />

of Princeton University<br />

in C<strong>and</strong>idacy <strong>for</strong> the Degree<br />

of Doctor of Philosophy<br />

Recommended <strong>for</strong> Acceptance<br />

by the Department of<br />

Civil <strong>and</strong> Environmental Engineering<br />

Advisor: Erik Vanmarcke<br />

April 2011


c○ Copyright by Siu Chung Yau, 2011.<br />

All Rights Reserved


Abstract<br />

Rising hurricane damages over the past half century have served to motivate policy makers<br />

<strong>and</strong> insurance industry executives to dem<strong>and</strong> better quantification of risk exposure to ex-<br />

treme winds. Advances in wind engineering <strong>and</strong> weather observation have enabled a shift<br />

away from empirically-based estimates, derived from insurance claim records, to more real-<br />

istic physically-based structural damage simulations. In this dissertation, the risk <strong>and</strong> reli-<br />

ability of low-rise structures under extreme winds are investigated using an advanced prob-<br />

abilistic scheme, a windborne debris model, <strong>and</strong> a novel pressure-based structural damage<br />

model. A review of several loss estimation methodologies, their sensitivities <strong>and</strong> inter-model<br />

discrepancies is provided. Specifically, the studies examined structural damage processes<br />

in three aspects: (1) structural components, (2) low-rise structures in isolation, <strong>and</strong> (3)<br />

neighborhoods of residences. Damage to the structural components was found to increase<br />

if the pressure acting on the components was either highly correlated with wind direction<br />

or wind direction varied over a larger range over time. The cumulative effect of internal<br />

pressure on the building envelope was analyzed using three methods: American Society of<br />

Civil Engineers 7 St<strong>and</strong>ard, Eurocode, <strong>and</strong> Florida Public Hurricane Loss Projection Model<br />

(FPHLPM). The percentage damage of the building envelope differed by a few percent to<br />

a few hundred percent depending on the wind speed, <strong>and</strong> structural configuration. The<br />

most striking results came from the application of a novel vulnerability model, which <strong>for</strong><br />

the first time, parameterized the interplay between pressure damage <strong>and</strong> windborne debris.<br />

Simulations <strong>for</strong>ced with probabilistic data from FPHLPM as well as wind-tunnel experi-<br />

ments revealed how a structure may be effected by the resistance of <strong>and</strong> the proximity to<br />

neighboring buildings. The vulnerability results were applied in a long-term risk assess-<br />

ment to illustrate the possible effects of a changing climate on structural life-cycle cost.<br />

Collectively, the ideas <strong>and</strong> model improvements presented in this dissertation represent a<br />

significant contribution to the development of the next generation of wind loss models.<br />

iii


Acknowledgements<br />

First of all, I owe my deepest gratitude to my advisor, Professor Erik Vanmarcke, <strong>for</strong> his<br />

guidance <strong>and</strong> support throughout my graduate study at Princeton. This thesis would not<br />

have been possible without his encouragement, insight <strong>and</strong> patience. It is truly my honor<br />

to be his student, learning from him about risk analysis, scientific field, <strong>and</strong> even different<br />

global issues. I would also like to thank the other thesis committee members, Professor<br />

Branko Gliˇsić, Professor Elie Bou-Zeid <strong>and</strong> Professor Maria Garlock, <strong>for</strong> their valuable<br />

suggestions <strong>and</strong> insightful comments.<br />

In addition, I gratefully acknowledge financial support <strong>for</strong> this research, in part through<br />

a grant to Princeton University from Baseline <strong>Management</strong> Company <strong>and</strong> in part under a<br />

project entitled “Improved Hurricane <strong>Risk</strong> <strong>Assessment</strong> with Links to Earth System Models,”<br />

funded through the Cooperative Institute <strong>for</strong> Climate Science (CICS) by the U.S. National<br />

Oceanographic <strong>and</strong> Atmospheric Administration (NOAA). Thanks are also due to Princeton<br />

Environmental Institute (PEI) <strong>for</strong> granting me the Program in Science, Technology, <strong>and</strong><br />

Environmental Policy (PEI-STEP) Fellowship.<br />

Next, I would like to express my gratitude to Professor Michael Oppenheimer <strong>and</strong> Pro-<br />

fessor Gregory van der Vink, who serve as my PEI-STEP fellowship advisors. I also need<br />

to give special thanks to Professor Henri Gavin at Duke University. Without his inspi-<br />

ration <strong>and</strong> guidance during my undergraduate study, I would not have chosen to pursue<br />

my doctoral degree at Princeton University. I am also thankful to Dr. Ning Lin <strong>for</strong> her<br />

collaboration.<br />

I am very <strong>for</strong>tunate to have many friends to accompany me throughout my days at<br />

Princeton. Dr. Craig Ferguson, who was my roommate <strong>for</strong> more than two years, has in-<br />

troduced me to American cooking <strong>and</strong> the exciting outdoor life. Serdar Selamet, my close<br />

friend since my undergraduate, has been the single person whom I see most <strong>and</strong> talk most to<br />

in these four <strong>and</strong> a half years. Raghav Pant, who shares numerous similar viewpoints with<br />

iv


me, has given me countless fun time. Yee Lok Wong has provided advice on mathematics<br />

as well as companionship remotely from Boston. I would also like to thank Joshua Fink,<br />

Eric Hui, Dr. Chitsomanus Muneepeerakul, Dr. Rachata Muneepeerakul, Zhihua Wang <strong>and</strong><br />

Dr. Yan Zhang <strong>for</strong> their company <strong>and</strong> many enjoyable conversations.<br />

Finally, I express my sincere gratitude to my family. Thanks to my parents, Richard<br />

<strong>and</strong> Elaine, <strong>and</strong> brother, Robee, <strong>for</strong> their care <strong>and</strong> encouragement during the preparation<br />

of this dissertation. Final thanks are due to my girlfriend, Chloe Kung, <strong>for</strong> her seemingly<br />

unending patience <strong>and</strong> support in every difficult moment of mine.<br />

v


Contents<br />

Abstract . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iii<br />

Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iv<br />

Chapter 1:<br />

Introduction 1<br />

1.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1<br />

1.2 Review on <strong>Wind</strong>-Related Vulnerability Models . . . . . . . . . . . . . . . . 2<br />

1.2.1 Full-Scale Extreme-<strong>Wind</strong> Loss Model . . . . . . . . . . . . . . . . . . . 2<br />

1.2.2 Qualitative Vulnerability Model . . . . . . . . . . . . . . . . . . . . . . 4<br />

1.2.3 Quantitative Vulnerability Model . . . . . . . . . . . . . . . . . . . . . 6<br />

1.2.4 Recent Developments . . . . . . . . . . . . . . . . . . . . . . . . . . . 9<br />

1.3 Objectives <strong>and</strong> Scope of Study . . . . . . . . . . . . . . . . . . . . . . . . . 10<br />

1.4 Organization of Dissertation . . . . . . . . . . . . . . . . . . . . . . . . . . . 12<br />

Chapter 2:<br />

Structural Components <strong>and</strong> <strong>Wind</strong> Directionality 13<br />

2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13<br />

2.2 Tropical-Cyclone <strong>Wind</strong> Time Series Model . . . . . . . . . . . . . . . . . . . 16<br />

2.2.1 Scope . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16<br />

2.2.2 Tropical Cyclone Properties . . . . . . . . . . . . . . . . . . . . . . . . 16<br />

2.2.3 Tropical Cyclone Track . . . . . . . . . . . . . . . . . . . . . . . . . . . 18<br />

vi


CONTENTS vii<br />

2.2.4 Mean Boundary Layer <strong>Wind</strong> . . . . . . . . . . . . . . . . . . . . . . . . 19<br />

2.2.5 Three-Second Surface <strong>Wind</strong> . . . . . . . . . . . . . . . . . . . . . . . . 21<br />

2.2.6 Damage Threshold . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22<br />

2.2.7 Sample Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23<br />

2.3 Alternative Approaches to <strong>Wind</strong>-Load <strong>and</strong> Reliability <strong>Assessment</strong> . . . . . 24<br />

2.3.1 <strong>Wind</strong> Load <strong>and</strong> Reliability Calculation . . . . . . . . . . . . . . . . . . 24<br />

2.3.2 Time-Stepping Method <strong>and</strong> Point-in-Time Method . . . . . . . . . . . . 25<br />

2.4 Numerical Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26<br />

2.4.1 Setup . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26<br />

2.4.2 Maximum <strong>Wind</strong> Pressure . . . . . . . . . . . . . . . . . . . . . . . . . 27<br />

2.4.3 Probability of Failure . . . . . . . . . . . . . . . . . . . . . . . . . . . 32<br />

2.5 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37<br />

2.6 Closing Remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39<br />

Chapter 3:<br />

Internal Pressure in Low-Rise <strong>Structures</strong> 40<br />

3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40<br />

3.2 Structural Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43<br />

3.3 External <strong>Wind</strong> Pressure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44<br />

3.4 Internal Pressure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45<br />

3.4.1 Florida Public Hurricane Loss Projection (FPHLP) Model . . . . . . . . 47<br />

3.4.2 ASCE 7 St<strong>and</strong>ard . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48<br />

3.4.3 Eurocode . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49<br />

3.5 <strong>Wind</strong>borne Debris Impact . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50<br />

3.6 Damage Evaluation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51<br />

3.6.1 Component Failure Probability . . . . . . . . . . . . . . . . . . . . . . 51<br />

3.6.2 System Failure Probability . . . . . . . . . . . . . . . . . . . . . . . . . 52


CONTENTS viii<br />

3.7 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54<br />

3.7.1 Baseline structure B under 150-mph wind . . . . . . . . . . . . . . . . . 54<br />

3.7.2 Roof Damage over 50–250 mph <strong>Wind</strong> Speed Range . . . . . . . . . . . . 59<br />

3.8 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67<br />

3.9 Closing Remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68<br />

Chapter 4:<br />

Integrated Multi-Structural Vulnerability Model 69<br />

4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69<br />

4.2 Methodology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70<br />

4.2.1 <strong>Wind</strong> Characteristics . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70<br />

4.2.2 Structural Characteristics . . . . . . . . . . . . . . . . . . . . . . . . . 72<br />

4.2.3 Component-Based Pressure Damage Model . . . . . . . . . . . . . . . . 75<br />

4.2.4 <strong>Wind</strong>borne Debris Model . . . . . . . . . . . . . . . . . . . . . . . . . 77<br />

4.2.5 Integrated Structural Damage Estimation . . . . . . . . . . . . . . . . . 79<br />

4.2.6 Approximate Economic Model . . . . . . . . . . . . . . . . . . . . . . . 81<br />

4.3 Numerical Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83<br />

4.3.1 A Hip-roof Concrete Residential Structure in Isolation . . . . . . . . . . . 83<br />

4.3.2 Hypothetical Residential Community . . . . . . . . . . . . . . . . . . . 83<br />

4.3.3 Two Neighboring Houses . . . . . . . . . . . . . . . . . . . . . . . . . . 88<br />

4.3.4 Sarasota, Florida . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89<br />

4.4 Closing Remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93<br />

Chapter 5:<br />

Long-Term <strong>Risk</strong> <strong>Assessment</strong> <strong>and</strong> <strong>Management</strong> 94<br />

5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94<br />

5.2 Methodology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95


CONTENTS ix<br />

5.3 <strong>Hazard</strong> model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96<br />

5.3.1 <strong>Wind</strong> Speed Distribution . . . . . . . . . . . . . . . . . . . . . . . . . . 96<br />

5.3.2 Compound Poisson Process . . . . . . . . . . . . . . . . . . . . . . . . 98<br />

5.4 Single-Limit-State Vulnerability Model . . . . . . . . . . . . . . . . . . . . . 102<br />

5.4.1 Model Description . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102<br />

5.4.2 Life-cycle Cost Results . . . . . . . . . . . . . . . . . . . . . . . . . . . 103<br />

5.4.3 Life-cycle-cost-related Decisions . . . . . . . . . . . . . . . . . . . . . . 109<br />

5.5 Modified Florida Public Hurricane Loss Projection (FPHLP) Model . . . . 113<br />

5.5.1 Model Description . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113<br />

5.5.2 Expected Repair Cost over the Contiguous United States . . . . . . . . . 116<br />

5.5.3 Expected Repair Cost <strong>for</strong> Different <strong>Structures</strong> . . . . . . . . . . . . . . . 118<br />

5.6 Closing Remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 120<br />

Chapter 6:<br />

Conclusions <strong>and</strong> Future Work 123<br />

6.1 Summary <strong>and</strong> Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123<br />

6.2 Future Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125<br />

6.2.1 Structural Components . . . . . . . . . . . . . . . . . . . . . . . . . . 125<br />

6.2.2 Low-Rise <strong>Structures</strong> . . . . . . . . . . . . . . . . . . . . . . . . . . . . 126<br />

6.2.3 Multi-Structural Communities . . . . . . . . . . . . . . . . . . . . . . . 127<br />

6.2.4 Low-Term <strong>Risk</strong> <strong>Assessment</strong> . . . . . . . . . . . . . . . . . . . . . . . . 128<br />

Bibliography 129


List of Figures<br />

2.1 Flowchart illustrating the tropical cyclone wind time series model <strong>and</strong> the<br />

difference between point-in-time <strong>and</strong> time-stepping methods. . . . . . . . . 15<br />

2.2 Overview of the tropical cyclone wind time series model. . . . . . . . . . . . 17<br />

2.3 Sample simulated wind speed time series <strong>and</strong> wind angle time series. . . . . 21<br />

2.4 <strong>Wind</strong> speed <strong>and</strong> wind angle time records during Hurricane Katrina <strong>and</strong> Wilma. 23<br />

2.5 Four selected pressure coefficients with respect to wind angle. . . . . . . . . 27<br />

2.6 Cumulative distributions of the maximum wind load <strong>and</strong> the approximated<br />

maximum wind load generated from the simulated wind time series <strong>for</strong> cat-<br />

egory 5 tropical cyclones. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29<br />

2.7 Kolmogorov-Smirnov statistics comparing the maximum wind load <strong>and</strong> the<br />

approximated maximum wind load. . . . . . . . . . . . . . . . . . . . . . . . 30<br />

2.8 Percentage differences between the probabilities of failure <strong>and</strong> the approxi-<br />

mated probabilities of failure <strong>for</strong> normally distributed resistances. . . . . . . 35<br />

2.9 Percentage differences between the probabilities of failure <strong>and</strong> the approxi-<br />

mated probabilities of failure <strong>for</strong> lognormally distributed resistances. . . . . 36<br />

3.1 Unfolded view of baseline structures. . . . . . . . . . . . . . . . . . . . . . . 42<br />

3.2 Component <strong>and</strong> cladding external pressure zones. . . . . . . . . . . . . . . . 46<br />

3.3 Flowchart of system damage simulation. . . . . . . . . . . . . . . . . . . . . 53<br />

x


LIST OF FIGURES xi<br />

3.4 Distributions of internal pressure coefficient <strong>and</strong> building envelope damage<br />

of baseline structure B under a 150-mph wind towards the structure’s front<br />

face. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55<br />

3.5 Distributions of internal pressure coefficient <strong>and</strong> building envelope damage<br />

of baseline structure B under a 150-mph wind towards the structure’s front<br />

face. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58<br />

3.6 Roof panel vulnerability curves. . . . . . . . . . . . . . . . . . . . . . . . . . 61<br />

3.7 Maximum percentage increase in mean roof damage. . . . . . . . . . . . . . 63<br />

3.8 Percentage differences between the average roof damages over four cardinal<br />

wind directions <strong>and</strong> the average roof damage of baseline structure C without<br />

windborne debris. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66<br />

4.1 Eight nominal angles of wind incidence. . . . . . . . . . . . . . . . . . . . . 71<br />

4.2 Three dimensional rendering of residential structure models. . . . . . . . . . 74<br />

4.3 Flowchart of the component-based pressure damage model. . . . . . . . . . 76<br />

4.4 Flowchart of the integrated vulnerability model. . . . . . . . . . . . . . . . . 80<br />

4.5 Vulnerability of a hip-roof concrete house at four different wind speeds. . . 84<br />

4.6 Vulnerability of a hip-roof concrete house under 200-mph at three different<br />

wind directions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85<br />

4.7 Three-dimensional view of a hypothetical residential development. . . . . . 86<br />

4.8 Estimated damage condition of each individual house in the hypothetical<br />

residential development under a wind speed of 65 m/s at a 45-degree angle. 87<br />

4.9 Comparison of subassembly <strong>and</strong> overall damage ratios with <strong>and</strong> without con-<br />

sideration of windborne debris. . . . . . . . . . . . . . . . . . . . . . . . . . 88<br />

4.10 (a) Illustration of two neighboring houses. (b) Breakdown of overall damage<br />

ratio of structure A versus resistance factor of structure B. . . . . . . . . . 89


LIST OF FIGURES xii<br />

4.11 Study area of a residential development of 358 houses in Sarasota County,<br />

Florida <strong>and</strong> simulated storm track of Hurricane Charley of 2004. . . . . . . 90<br />

4.12 Comparison of the estimated wind damage condition of the residential de-<br />

velopment in Sarasota County, Florida, under a Hurricane Charley wind<br />

condition, by time series <strong>and</strong> maximum wind analyses. . . . . . . . . . . . . 91<br />

4.13 Comparison of subassembly <strong>and</strong> overall damage ratios between time-series<br />

<strong>and</strong> maximum-wind damage analyses. . . . . . . . . . . . . . . . . . . . . . 92<br />

5.1 Distribution of Gumbel distribution parameters λ <strong>and</strong> µ of extreme wind<br />

speed data in 129 stations over the contiguous United States. . . . . . . . . 100<br />

5.2 Two sample results of the hazard model. . . . . . . . . . . . . . . . . . . . . 101<br />

5.3 Vulnerability curves of the single-limit-state deterministic model. . . . . . . 103<br />

5.4 Mean total discounted repair cost in 36 different climate change scenarios. . 104<br />

5.5 Percentage difference in mean total discounted repair costs between 36 dif-<br />

ferent climate change scenarios <strong>and</strong> the ‘no climate change’ scenario. . . . . 105<br />

5.6 Mean total discounted repair cost in 36 different climate change scenarios. . 107<br />

5.7 Percentage difference in mean total discounted repair costs between 36 dif-<br />

ferent climate change scenarios <strong>and</strong> the ‘no climate change’ scenario. . . . . 108<br />

5.8 Ratios of the total mean discounted repair costs attributed to the uncertainty<br />

in parameters to that attributed to the mean value of parameters. . . . . . 110<br />

5.9 Ratios of the st<strong>and</strong>ard deviations of the discounted repair cost attributed<br />

to the uncertainty in parameters to that attributed to the mean value of<br />

parameters. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110<br />

5.10 Four sample simulations showing the effects of resistance deterioration in two<br />

different climate scenarios. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112<br />

5.11 Minimization of life-cycle cost in two different climate scenarios. . . . . . . 114


LIST OF FIGURES xiii<br />

5.12 Examples of vulnerability curves of the modified Florida Public Hurricane<br />

Loss Model. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116<br />

5.13 Expected total repair cost of a one-story hip-roof concrete-block house over<br />

129 geographic locations in contiguous United States in nine different future<br />

climate scenarios. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117<br />

5.14 Expected total repair cost related to structural damage: (a) hip roof vs.<br />

gable roof (b) tiles vs. shingles (c) shutters vs. no shutters. . . . . . . . . . 119<br />

5.15 Vulnerability curves <strong>and</strong> total expected repair cost related to structural dam-<br />

age of a concrete-block house at four different geographic locations. . . . . . 120<br />

5.16 Vulnerability curves <strong>and</strong> total expected repair cost related to structural dam-<br />

age of a wood-frame house at four different geographic locations. . . . . . . 121


List of Tables<br />

1.1 The five costliest U.S. hurricanes, 1900-2006. . . . . . . . . . . . . . . . . . 2<br />

1.2 Saffir-Simpson hurricane damage potential scale. . . . . . . . . . . . . . . . 4<br />

1.3 Building classification by Minor <strong>and</strong> Mehta (1979). . . . . . . . . . . . . . . 5<br />

2.1 Probability distribution of tropical cyclone parameters. . . . . . . . . . . . . 18<br />

2.2 Percentage of simulated wind records by track location. . . . . . . . . . . . 32<br />

2.3 Probabilities of failure of four component limit states . . . . . . . . . . . . . 34<br />

3.1 Statistics of structural component resistance. . . . . . . . . . . . . . . . . . 44<br />

3.2 Statistics of external pressure coefficient. . . . . . . . . . . . . . . . . . . . . 46<br />

4.1 Example parameters of a residential structure’s geometry. . . . . . . . . . . 73<br />

4.2 Limit states of structural components. . . . . . . . . . . . . . . . . . . . . . 75<br />

4.3 Properties of debris objects. . . . . . . . . . . . . . . . . . . . . . . . . . . . 78<br />

5.1 Expected total discounted repair costs <strong>and</strong> maximum maintenance cost in<br />

two different climate change scenarios. . . . . . . . . . . . . . . . . . . . . . 111<br />

5.2 Gumbel distribution parameters of four different geographic locations with<br />

similar 50-year wind speed. . . . . . . . . . . . . . . . . . . . . . . . . . . . 119<br />

xiv


1.1 Overview<br />

Chapter<br />

1<br />

Introduction<br />

Death <strong>and</strong> injuries in windstorms are relatively low compared to the casualty rates in other<br />

natural disasters, due to the predictability of extreme winds. However, the relative low<br />

number of casualties does not sufficiently reflect the massive destructiveness of extreme<br />

winds. <strong>Wind</strong>storm (including tropical cyclone, tornado <strong>and</strong> thunderstorm) is the most<br />

costly type of natural disaster in the United states in recent decades (EM-DAT, 2010).<br />

Hurricane Andrew, which caused an insured loss of $16.5 billion in 1992, deeply impacted<br />

the insurance industry. Twelve insurers bankrupted <strong>and</strong> most of the catastrophe capital<br />

in the world reinsurance market is depleted (Malmquist <strong>and</strong> Michaels, 2000). In 2005,<br />

Hurricane Katrina caused $81 billion of property damage (Blake et al., 2007) <strong>and</strong> $40.6<br />

billion of insured loss (Knabb et al., 2005). Table 1.1 (Blake et al., 2007) lists the five<br />

costliest United States hurricanes.<br />

In light of the recent massive amounts of damage caused by tropical cyclones, wind-<br />

related vulnerability of structures has received increasing attention in the research field.<br />

Traditionally, structural damage caused by extreme winds is approximated by some em-<br />

1


Chapter 1. Introduction 2<br />

Table 1.1: The five costliest U.S. hurricanes, 1900-2006 (Blake et al., 2007).<br />

Hurricane Date Cost of Damage (2006 dollar value)<br />

Katrina August 2005 $84.6 billion<br />

Andrew August 1992 $48.1 billion<br />

Wilma October 2005 $21.5 billion<br />

Charley August 2005 $16.3 billion<br />

Ivan September 2004 $15.5 billion<br />

pirical relationships between wind speed <strong>and</strong> insurance-claim amounts determined from<br />

historical data (e.g. Ch<strong>and</strong>ler et al., 2001; Huang et al., 2001; Sparks, 2003). However,<br />

such relationships may be inaccurate because insurance records are strongly influenced by<br />

human factors (Watson <strong>and</strong> Johnson, 2004). They also do not predict damage to build-<br />

ings that have not been subjected to major wind hazards (Watson <strong>and</strong> Johnson, 2004).<br />

There<strong>for</strong>e, some recent studies have employed structural engineering analysis to simulate<br />

the actual physical damage processes <strong>and</strong> hence the economic loss due to structural dam-<br />

age (e.g. Stubbs <strong>and</strong> Perry, 1996; Pinelli et al., 2004; Unanwa et al., 2000; Vickery et al.,<br />

2006b). Per<strong>for</strong>mance-based design, which is based on the structural damage simulation, was<br />

proposed to strengthen future residences’ resistance against high winds (e.g. van de Lindt<br />

<strong>and</strong> Dao, 2009). By developing a more accurate <strong>and</strong> in-depth structural risk assessment<br />

in tropical cyclone-prone areas, engineers can help the actuaries <strong>and</strong> policymakers better<br />

prepare <strong>for</strong> <strong>and</strong> mitigate the future risk of extreme wind posed to our communities.<br />

1.2 Review on <strong>Wind</strong>-Related Vulnerability Models<br />

1.2.1 Full-Scale Extreme-<strong>Wind</strong> Loss Model<br />

A full-scale extreme-wind loss estimation model includes, but not limited to, calculation of<br />

genesis of extreme winds, evaluation of wind effects on buildings, estimation of insurance<br />

claim amounts based on building damage, etc ˙ , which require knowledge in a range of pro-


Chapter 1. Introduction 3<br />

fessions including meteorology, wind engineering, structural engineering, actuarial science.<br />

In general, the model may be divided into four modules:<br />

1. <strong>Wind</strong> Model<br />

Given the historical storm tracks <strong>and</strong> a l<strong>and</strong> cover model, a wind model determines<br />

the occurrence <strong>and</strong> intensity of wind hazard such as wind speed <strong>and</strong> wind direction.<br />

One commonly-used simple model <strong>for</strong> tropical cyclones is the Rankine Vortex wind<br />

field model.<br />

2. Boundary Layer Model<br />

The wind hazard in<strong>for</strong>mation generated by the large-scale wind field model is modified<br />

in a boundary layer model to take into account the effects of l<strong>and</strong> topography. In<strong>for</strong>-<br />

mation about the local wind field, which impacts the buildings <strong>and</strong> the environment,<br />

is obtained.<br />

3. Vulnerability Model<br />

A vulnerability model relates the extreme wind data to the consequent physical dam-<br />

age of buildings in a geographic area. The damage is expressed as the failure proba-<br />

bility of some structural components of the buildings, <strong>and</strong> may also be represented in<br />

monetary terms by the cost to repair the damage.<br />

4. Insurance Loss Model<br />

The economic losses are approximated based on the structural damage by an insurance<br />

loss model which considers the economy situation <strong>and</strong> insurance policy terms such as<br />

deductibles.<br />

The wind model <strong>and</strong> the boundary layer model are usually subjects of meteorology<br />

<strong>and</strong> aerodynamics, while the insurance loss model is the study of actuarial science. This<br />

dissertation, as well as the following literature review, mainly focuses on the vulnerability


Chapter 1. Introduction 4<br />

model from a structural engineering st<strong>and</strong>point. Most vulnerability models are about low-<br />

rise structures, particularly residential houses, since they are the most commonly damaged<br />

structural type in extreme wind events.<br />

1.2.2 Qualitative Vulnerability Model<br />

<strong>Wind</strong> damage to a building may be estimated via both qualitative <strong>and</strong> quantitative meth-<br />

ods. Qualitative vulnerability models categorize different building types or different wind<br />

events, <strong>and</strong> describe the expected damage level of buildings at a range of wind intensities<br />

(Unanwa et al., 2000). An example of qualitative vulnerability model is the Saffir-Simpson<br />

hurricane damage potential scale as shown in Table 1.2. Another example is the classifi-<br />

cation of buildings developed by Minor <strong>and</strong> Mehta (1979) presented in Table 1.3, which<br />

categorizes a building as fully engineered, pre-engineered, marginally engineered, or non-<br />

engineered, with their associated per<strong>for</strong>mance in extreme winds.<br />

Table 1.2: Saffir-Simpson hurricane damage potential scale.<br />

Category <strong>Wind</strong> Speed (mph) Damage Level<br />

1 74-95 Minimal<br />

2 96-110 Moderate<br />

3 111-130 Extensive<br />

4 131-155 Extreme<br />

5 >155 Catastrophic<br />

The categorizations used by the qualitative vulnerability models are developed <strong>for</strong> gen-<br />

eral purposes only. They may be useful notifying the general public what to expect <strong>for</strong><br />

different building types in various wind scenarios. However, they are not precise enough to<br />

predict hurricane damage or insurance loss to any specific buildings. Quantitative vulnera-<br />

bility models are necessary <strong>for</strong> accurate prediction of losses.


Chapter 1. Introduction 5<br />

Table 1.3: Building classification by Minor <strong>and</strong> Mehta (1979).<br />

Building Description Per<strong>for</strong>mance<br />

Fully-engineered<br />

buildings<br />

Pre-engineered<br />

buildings<br />

Marginally engineered<br />

buildings<br />

Non-engineered<br />

buildings<br />

Buildings that receive specific, individualized<br />

design attention from<br />

professional architects <strong>and</strong> engineers<br />

Buildings of this type receive engineering<br />

attention in advance of<br />

a commitment to constmction <strong>and</strong><br />

are subsequently marketed in many<br />

similar units <strong>for</strong> a wide variety of<br />

uses.<br />

Commercial buildings, light industrial<br />

buildings, schools, <strong>and</strong> certain<br />

types of motels <strong>and</strong> apartments<br />

built with some combination of masonry,<br />

light steel framing, openweb<br />

joists, wood framing, <strong>and</strong> wood<br />

rafters falls under this category.<br />

Single <strong>and</strong> multifamily residences,<br />

one or two story apartment units<br />

<strong>and</strong> many small commercial buildings<br />

receive no engineering attention<br />

at all <strong>and</strong> are called non-engineered<br />

Per<strong>for</strong>mance in the face of<br />

extreme winds is likely to<br />

be very good.<br />

Damages can occur at 125<br />

mph, if care is not exercised<br />

in design <strong>and</strong> construction.<br />

Total destruction may result<br />

at wind speed of 125<br />

mph or less.<br />

<strong>Wind</strong> speeds of 75 mph<br />

represent a threshold of<br />

damage, <strong>and</strong> total destruction<br />

may occur when winds<br />

reach 125 mph.


1.2.3 Quantitative Vulnerability Model<br />

Chapter 1. Introduction 6<br />

Generally, there are three different approaches to develope a quantitative vulnerability<br />

model (Watson <strong>and</strong> Johnson, 2004): (1) claim-based, (2) engineering judgment, <strong>and</strong> (3)<br />

theory-based. Each of the approaches has its own advantages over the others <strong>and</strong> is improved<br />

over time. Some vulnerability models employ more than one approach simultaneously.<br />

1.2.2.1 Claim-based Approach<br />

Vulnerability models using the claim-based approach estimate the hurricane damage of a<br />

building without considering the physics or engineering of the building components, <strong>and</strong><br />

usually derive relationships between the historical damage <strong>and</strong> wind speeds from the past<br />

insurance claim data. This approach may not predict the damage precisely mainly because<br />

the insurance loss in<strong>for</strong>mation, including the amount of claims <strong>and</strong> the date of claims, is<br />

largely controlled by human factors. Such factors may include personal view of an individual<br />

insurance adjuster <strong>and</strong> aggressiveness of homeowner (Watson <strong>and</strong> Johnson, 2004). For ex-<br />

ample, the amount a homeowner claims is affected by the policy details such as deductibles.<br />

The company policies regarding insurance claims are subjected to administrative <strong>and</strong> polit-<br />

ical considerations that differ from storm to storm. Moreover, the insurance claim data in<br />

a region may not apply to other regions. Factors including environment <strong>and</strong> construction<br />

practices vary from region to region.<br />

Many vulnerability models are developed using the claim-based approach. Studies by<br />

Leicester et al. (1979) <strong>and</strong> Leicester (1981) are those of the first studies in the literature that<br />

established <strong>and</strong> elaborated the concepts of vulnerability curves. In their models, Leicester<br />

et al. adopted a linear relationship between the damage index <strong>and</strong> the gust wind speed<br />

<strong>for</strong> different types of housing in Australia based on damage surveys. Contrast to the linear<br />

relationship between wind speed <strong>and</strong> damage, Sparks et al. (1994) suggested that significant<br />

damage begins when the gradient wind speed reaches 40 m/s <strong>and</strong> rises steadily until 70 m/s.


Chapter 1. Introduction 7<br />

Between 70 m/s <strong>and</strong> 80 m/s, there is a sudden rise in damage due to loss of building envelope<br />

<strong>and</strong> rain entering the building. Sparks et al. (1994) also observed that housing damage<br />

is closely related to the per<strong>for</strong>mance of the roof <strong>and</strong> wall building envelope. Similarly,<br />

Mitsuta et al. (1996) observed a highly correlated relationship between home damage <strong>and</strong><br />

surface wind speed from insurance data of tropical cyclones Mireille <strong>and</strong> Flo in Japan. Sill<br />

<strong>and</strong> Kozlowski (1997) developed an approach <strong>for</strong> predicting the hurricane damage <strong>for</strong> a<br />

population of structures as a function of a wide range of factors including gradient wind<br />

speed, gust factor, average value of the buildings, <strong>and</strong> two parameters which govern the<br />

rate of damage increase with wind speed. The proposed method was intended to move<br />

away <strong>for</strong> curve fitting schemes, but its practical value is hampered by insufficient clarity<br />

<strong>and</strong> transparency (Pinelli et al., 2004).<br />

1.2.2.2 Engineering judgment Approach<br />

The engineering judgment approach is based on the engineering survey of damaged struc-<br />

tures (Watson <strong>and</strong> Johnson, 2004). Similar to the claim-based approach, it largely depends<br />

on individual interpretation. For instance, there are many different versions about the<br />

definition of a 50% damage of a roof (Watson <strong>and</strong> Johnson, 2004).<br />

One example of the engineering judgment approaches is the analysis of data obtained<br />

from the American Red Cross data on residential damage from hurricanes by Howard et al.<br />

(1972). Also, engineering judgment often supplements the theory-based approach. For<br />

example, the Florida public hurricane loss projection model (Gurley et al., 2005b) uses<br />

engineering judgment to modify the ASCE 7 wind load specification <strong>for</strong> hurricane loss<br />

estimation.


1.2.2.3 Theory-based Approach<br />

Chapter 1. Introduction 8<br />

In the theory-based approach, the vulnerability of the structure is determined based on the<br />

physics of behaviors of the structure (Watson <strong>and</strong> Johnson, 2004). The approach explicitly<br />

characterizes the resistance of various building components <strong>and</strong> the wind load effects on<br />

them. Damage to a component occurs when the load effect is greater than its resistance.<br />

The vulnerability of a structure at various wind speed can be estimated once the strength<br />

capacities, load dem<strong>and</strong>s <strong>and</strong> load paths within a structure are identified (Pinelli et al.,<br />

2004). Unlike the claim-based approach, the theory-based approach is able to evaluate the<br />

hurricane vulnerability of structures without the heavy influence of human judgment.<br />

Some recent examples of the theory-based vulnerability models are by Chiu (1994),<br />

Holmes (1996), Stubbs (1996), Unanwa <strong>and</strong> McDonald (2000) <strong>and</strong> Unanwa et al. (2000).<br />

Holmes (1996) presented the vulnerability curve <strong>for</strong> a fully-engineered building with re-<br />

sistance of lognormal distribution. It was assumed that the failure of each component is<br />

independent of each other <strong>and</strong> is designed to resist the same value of wind load. The study<br />

indicates that more thorough post-disaster investigations are needed to better define the<br />

model. Stubbs (1996) proposed a vulnerability model which is based on an analysis of the<br />

per<strong>for</strong>mance of different building components <strong>and</strong> their corresponding relative importance.<br />

The damage ratio <strong>for</strong> the entire structure is weighted in proportion to some constants rep-<br />

resenting the relative importance of the contribution of the consequence of each damage<br />

mode. However, as indicated by Unanwa <strong>and</strong> McDonald (2000), this study is restricted to<br />

damage prediction of a group of buildings within a given geographic area. Unanwa <strong>and</strong><br />

McDonald (2000) developed a similar approach as Stubbs (1996), but their model is more<br />

versatile <strong>and</strong> more capable of predicting wind damage <strong>for</strong> a large number of different house<br />

types <strong>and</strong> building per<strong>for</strong>mance parameters.


1.2.4 Recent Developments<br />

Chapter 1. Introduction 9<br />

Damage analysis of a building can be very complex, since it involves a lot of issues <strong>and</strong><br />

uncertainties. For example, wind pressure acting on the roof largely depends on the roof<br />

shape, roof angle <strong>and</strong> even presence of nearby trees. Internal pressure inside a building is<br />

affected by overall building leakage, flexibility of the building envelope <strong>and</strong> compartmental-<br />

ization within the building. Other related issues range from the surrounding environment of<br />

the structure to the time effect on the building material’s strength. Many of these complex<br />

issues have been simplified by design codes <strong>and</strong> st<strong>and</strong>ards <strong>for</strong> structural design purpose.<br />

However, <strong>for</strong> the purpose of accurate loss estimation, currently there are still significant<br />

uncertainties regarding the wind damage process such as the wind-borne debris impacts<br />

<strong>and</strong> the progressive damage, leaving plenty of room <strong>for</strong> development in the literature.<br />

Most current models analyze the vulnerability of a structure individually without con-<br />

sidering the surrounding environment, especially other neighboring structures. Presence<br />

of neighboring structures not only affects the local wind intensity <strong>and</strong> direction, but also<br />

contributes to the hurricane debris that often causes severe structural damages. Hence,<br />

one of the recent directions is the effect of wind-borne debris on structures during tropical<br />

cyclones. An example of the recent developments is the work by Lin <strong>and</strong> Vanmarcke (2010).<br />

In addition, many vulnerability models estimate losses of buildings as it was during the time<br />

period when they were constructed. As a result, the hurricane losses of the aging buildings<br />

may not be estimated accurately. There<strong>for</strong>e, some recent research studies focus on modeling<br />

the time effect on vulnerability by incorporating changes in housing types, new materials,<br />

age profiles, building code specifications. For example, Stewart et al. (2003) used the vul-<br />

nerability model developed by Huang et al. (2001) <strong>and</strong> extended the method to include<br />

changes in existing residential structural vulnerability due to improvements in building en-<br />

velope per<strong>for</strong>mance. Davidson <strong>and</strong> Rivera (2003) introduced a quantitative methodology<br />

to model how the vulnerability of a regions building inventory changes over time due to,


Chapter 1. Introduction 10<br />

<strong>for</strong> example, aging, structural upgrading ef<strong>for</strong>ts, construction <strong>and</strong> demolition of buildings.<br />

Also, Kh<strong>and</strong>uri <strong>and</strong> Morrow (2003) presented a method to disaggregate a vulnerability curve<br />

into several curves representing different building types so that a known vulnerability curve<br />

from one region can be applied to another region of a different building inventory. Pinelli<br />

et al. (2004) proposed a probabilistic approach <strong>for</strong> evaluation of hurricane vulnerability of<br />

residential structures. The model first defines the basic damage modes <strong>for</strong> components of<br />

specific building types. Then the damage modes are combined into possible damage states,<br />

of which the probabilities of occurrence are calculated as functions of wind speeds from<br />

Monte Carlo simulations conducted on engineering numerical models of typical houses.<br />

The above recent developments are not realized in practice until they are applied to<br />

the loss estimation models that are used by the catastrophe modeling industry. While<br />

a number of proprietary commercial models are available to the insurance industry <strong>for</strong><br />

estimating the potential insurance loss caused by hurricanes, currently there are only a<br />

very limited number of hurricane loss estimation models accessible in the public domain.<br />

One of the most well-known public software is the HAZUS-MH model that is developed<br />

by the Federal Emergency <strong>Management</strong> Agency (FEMA). However, a lot of details <strong>and</strong><br />

methodologies implemented in the model are not disclosed to the public. Recently a new<br />

hurricane loss estimation model, namely Florida Public Hurricane Loss Projection (FPHLP)<br />

Model is approved by the Florida Commission on Hurricane Loss Projection Methodology<br />

<strong>for</strong> estimating hurricane insurance rates in the state of Florida. One of its goals is to provide<br />

a more publicly accessible model. However, he model is confined to analysis of structures<br />

in Florida, <strong>and</strong> is not fully open-source.<br />

1.3 Objectives <strong>and</strong> Scope of Study<br />

Very recently, the research field has started full-scale wind tunnel experiments, such as the<br />

‘Wall of <strong>Wind</strong>’ (WoW) facility at Florida International University, to simulate the damage


Chapter 1. Introduction 11<br />

of low-rise structures under tropical cyclone winds. As full-scale experiments may be costly<br />

in both time <strong>and</strong> money, it is important to identify the critical issues that can utilize<br />

the facilities. So far, many research results in the literature are obtained from numerical<br />

analysis. The first objective of this dissertation is to identify the discrepancies among<br />

different numerical methodologies in vulnerability analysis, indicating the issues that have<br />

to be studied in the physical experiments.<br />

Moreover, presently, the very limited number of open-source hurricane loss estimation<br />

models impedes the improvement of the methodology. The details <strong>and</strong> methodologies em-<br />

bedded in many current models are concealed from the public so that it is difficult to<br />

examine the difference in loss estimation results of different geographic regions, building<br />

types <strong>and</strong> details among different analysis methods. The second objective of this disserta-<br />

tion is to review the details of the loss estimation methodology <strong>and</strong> examine the effects of<br />

the details on the vulnerability analysis results.<br />

Thirdly, effect of climate change on structural engineering analysis has been addressed<br />

by very limited literature only. Optimization of structural design requires evaluation of<br />

long-term structural per<strong>for</strong>mance, which is sensitive to any change in future extreme wind<br />

load. The third objective of this dissertation is to study the potential impact of climate<br />

change on the long-term risk assessment <strong>and</strong> management of structures.<br />

<strong>Risk</strong> assessment of structures under extreme winds is a complex <strong>and</strong> sizable topic. As<br />

specified in the last section, this dissertation focuses on the vulnerability analysis of struc-<br />

tures, which relates the extreme wind data to the consequent physical damage of buildings.<br />

The genesis of extreme winds <strong>and</strong> related aerodynamics such as boundary layer modeling<br />

are not considered. Also, this dissertation does not look into the details of the economic<br />

loss estimation, particularly the insurance policies of damaged structures. In addition, the<br />

structural analysis is very different <strong>for</strong> high-rise <strong>and</strong> low-rise structures. Under tropical cy-<br />

clone winds, high-rise structures are subjected to dynamic loading while low-rise structures


Chapter 1. Introduction 12<br />

are subjected to quasi-static wind load. A majority of this dissertation is devoted to the<br />

estimation of the structural damage of one-story residential structure.<br />

1.4 Organization of Dissertation<br />

This dissertation is constituted of two parts. The first part studies the vulnerability of resi-<br />

dential structures in individual wind events, starting from building components to structural<br />

systems <strong>and</strong> residential communities. First, the effect of wind directionality on the vulner-<br />

ability of building components is studied (Chapter 2). Using a numerical tropical cyclone<br />

wind time series model, the study contrasts the difference in damage predictions between<br />

time series analysis <strong>and</strong> point-in-time analysis during a storm passage. Then, the interac-<br />

tion between different building components is investigated in the risk assessment of low-rise<br />

structures (Chapter 3). Different internal pressure adjustment methods <strong>and</strong> their effects<br />

on the loss estimation are compared. The last module presents an integrated vulnerability<br />

model that considers the interplay between pressure damage <strong>and</strong> wind-borne debris impact,<br />

<strong>and</strong> applies the model to the damage estimation of a community of residential structures<br />

(Chapter 4). The second part of the dissertation moves the focus from individual wind<br />

events to the long term risk assessment of structures. The results of structural vulner-<br />

ability in individual wind events are applied to estimation of life-cycle cost of structures<br />

under different climate change scenarios (Chapter 5). A non-stationary stochastic pro-<br />

cess is developed to simulate the long-term wind climate. The wind results are applied to<br />

two vulnerability models to illustrate the possible effects of climate change on the mean<br />

<strong>and</strong> st<strong>and</strong>ard deviation of future repair cost. The end of the dissertation summarizes the<br />

findings, draws conclusions, <strong>and</strong> outlines areas of future work (Chapter 6).


Chapter<br />

2<br />

Structural Components <strong>and</strong><br />

<strong>Wind</strong> Directionality<br />

2.1 Introduction<br />

This chapter focuses on the effect of varying wind directionality on structural damage in<br />

individual tropical cyclone events. During a tropical cyclone event, a structure is exposed<br />

to winds from different directions as the cyclone moves along its track. The greater the<br />

variation in wind directionality, the wider the structure’s angle of exposure to windward<br />

wind. (“<strong>Wind</strong>ward” is the direction from which the wind is blowing at a given time, while<br />

“leeward” is the direction downwind from the point of reference.) If windward <strong>and</strong> leeward<br />

winds cause different types or levels of structural damage, this may significantly affect the<br />

damage estimation. For instance, more damage can be caused by wind-borne debris if wind<br />

directionality changes more, as debris mainly impacts the windward side of the structure<br />

(Lin <strong>and</strong> Vanmarcke, 2010). Moreover, failures of structural components are interdependent.<br />

After the initial structural damage, a change in wind direction can significantly affect the<br />

subsequent occurrences of damage. This effect of interdependent component failures cannot<br />

be quantified reliably without considering the change in wind speed <strong>and</strong> wind direction over<br />

13


Chapter 2. Structural Components <strong>and</strong> <strong>Wind</strong> Directionality 14<br />

time within individual strong-wind events.<br />

Currently, two methods are commonly used in different tropical cyclone damage esti-<br />

mation models. The first considers the effect of varying wind directionality during a strong<br />

wind event <strong>and</strong> the second does not. In the first, the time-stepping method, employed in<br />

damage estimation models such as the HAZUS-MH model, the structural loading is eval-<br />

uated at a series of time steps, e.g. at 10-minute interval, <strong>for</strong> both wind speed <strong>and</strong> wind<br />

direction (Vickery et al., 2006a,b). The second method, the point-in-time method, used<br />

in other models, including the Florida Public Hurricane Loss Projection Model, evaluates<br />

the wind load only at the maximum wind speed during the tropical cyclone event (<strong>and</strong><br />

the corresponding wind direction at the maximum wind speed) (Gurley et al., 2006; Li<br />

<strong>and</strong> Ellingwood, 2006). If the two methods are applied to the same quasi-static wind load<br />

model (used by both HAZUS-MH <strong>and</strong> the Florida Public Hurricane Loss Projection Model,<br />

since fatigue <strong>and</strong> dynamic loading are generally not considered), the difference in results is<br />

attributable to the variation in wind direction over time.<br />

The flowchart in Figure 2.1 summarizes the simplified loss estimation framework adopted<br />

in this study. In what follows, we examine <strong>and</strong> quantify the effect of varying wind direc-<br />

tionality on low-rise structural damage estimation, illustrating the methodology through<br />

numerical examples. In the first part of this chapter, we present an efficient probabilistic<br />

model to generate wind speed <strong>and</strong> wind direction time series in tropical cyclones. These<br />

are used, in the second part, to calculate wind loads <strong>and</strong> analyze the reliability of some<br />

representative structural components of low-rise structures.


Chapter 2. Structural Components <strong>and</strong> <strong>Wind</strong> Directionality 15<br />

Maximum <strong>Wind</strong> Speed (Vmax)<br />

Approximated Max <strong>Wind</strong> Load (S*)<br />

Tropical Cyclone Properties (Vt, !P, Rmax)<br />

Track Location (s)<br />

Tropical Cyclone <strong>Wind</strong> Time Series Model<br />

<strong>Wind</strong> Speed Time Series (V(t))<br />

<strong>Wind</strong> Angle Time Series (!(t))<br />

Point-in-Time Method Time-Stepping Method<br />

Pressure Coefficient<br />

Reliability Analysis<br />

over all t<br />

Pressure Coefficient<br />

Maximum <strong>Wind</strong> Load (Smax)<br />

Reliability Analysis<br />

Approximated Probability of Failure (Pf*) Probability of Failure (Pf)<br />

Figure 2.1: Flowchart illustrating the tropical cyclone wind time series model <strong>and</strong> the<br />

difference between point-in-time <strong>and</strong> time-stepping methods.


Chapter 2. Structural Components <strong>and</strong> <strong>Wind</strong> Directionality 16<br />

2.2 Tropical-Cyclone <strong>Wind</strong> Time Series Model<br />

2.2.1 Scope<br />

Due to the lack of historical tropical cyclone records, many advanced meteorological models<br />

have been developed to simulate wind data <strong>for</strong> various geographical regions (e.g. Powell<br />

et al., 2005; Vickery et al., 2000). They tend to be computationally expensive to solve, <strong>and</strong><br />

their solutions, typically in the <strong>for</strong>m of a two-dimensional time-dependent field, provide<br />

much more in<strong>for</strong>mation than is needed <strong>for</strong> the purpose of structural damage estimation, in<br />

particular in this study. A geographically non-specific parametric model that reasonably<br />

approximates the results of the wind field equations at a single location is judged most<br />

suitable <strong>for</strong> use in this study. Figure 2.2 indicates the methodology involved.<br />

2.2.2 Tropical Cyclone Properties<br />

First, the intensity <strong>and</strong> the spatial scale of the tropical cyclone are sampled. These are<br />

measured by three parameters: (1) the translational speed of the storm, Vt, namely the<br />

speed at which the tropical cyclone is moving along the earth surface; (2) the central<br />

pressure difference, ∆P , being the difference between atmospheric pressures at the center<br />

of the storm <strong>and</strong> at the periphery; <strong>and</strong> (3) the radius of maximum winds, Rmax, which is<br />

the distance from the storm center to the locations of maximum wind speed.<br />

Models have been suggested <strong>for</strong> the probability distributions of these parameters (Geor-<br />

giou 1985; Vickery <strong>and</strong> Twisdale 1995a). In this study, we adopt the simple probability<br />

distributions used by Iman et al. (2002), which are not specific to any geographic region;<br />

each of the three parameters (Vt, ∆P <strong>and</strong> Rmax) is characterized by a triangular distri-<br />

bution <strong>and</strong> the parameters are assumed to be statistically independent. The distributions<br />

are defined <strong>for</strong> each of the Saffir-Simpson hurricane categories 1, 3 <strong>and</strong> 5 in the study by<br />

Iman et al. (2002), according to which the parameters of the probability distributions are


(deg)<br />

V (m/s)<br />

60<br />

40<br />

20<br />

0<br />

180<br />

90<br />

0<br />

90<br />

Storm<br />

Track<br />

Chapter 2. Structural Components <strong>and</strong> <strong>Wind</strong> Directionality 17<br />

Structure<br />

Location<br />

SETUP<br />

s d(t)<br />

t (hr)<br />

Storm Center<br />

Position at t<br />

!(t)<br />

180<br />

0 5 10 15 20<br />

t (hr)<br />

d (km)<br />

(deg)<br />

V env (m/s)<br />

V tan (m/s)<br />

V rad (m/s)<br />

RELATIVE POSITION BETWEEN STORM CENTER<br />

AND STRUCTURE<br />

400<br />

200<br />

0<br />

180<br />

90<br />

0<br />

90<br />

t (hr)<br />

180<br />

0 5 10 15 20<br />

t (hr)<br />

WIND VELOCITY TIME SERIES WIND COMPONENTS<br />

50<br />

0<br />

50<br />

50<br />

0<br />

50<br />

50<br />

0<br />

5 10 15<br />

t (hr)<br />

5 10 15<br />

t (hr)<br />

50<br />

0 5 10<br />

t (hr)<br />

15 20<br />

Figure 2.2: Overview of the tropical cyclone wind time series model. The two upper subfigures<br />

illustrate the methodology of determining distance d(t) <strong>and</strong> relative angle φ(t) between<br />

the structure <strong>and</strong> the storm center over time. Based on d(t) <strong>and</strong> φ(t), the three tropical<br />

cyclone wind components: environmental wind (Venv), storm-scale tangential wind (Vtan)<br />

<strong>and</strong> storm-scale radial wind (Vrad) are computed, as shown in the lower right subfigure. The<br />

sum of these components is the wind velocity timer series in the lower left subfigure.


Chapter 2. Structural Components <strong>and</strong> <strong>Wind</strong> Directionality 18<br />

determined by a professional team in the Florida Commission on Hurricane Loss Projection<br />

Methodology. For example, the statistical measures (lower limit, mode <strong>and</strong> upper limit) of<br />

Vt, ∆P <strong>and</strong> Rmax <strong>for</strong> a category 5 hurricane are (4.5, 6.75, 9) (85, 95, 105) <strong>and</strong> (8, 20, 40),<br />

respectively.<br />

Table 2.1: Probability distribution of tropical cyclone parameters.<br />

Parameter Symbol Unit Category 1 Category 3 Category 5<br />

Translational velocity Vt m/s<br />

Central pressure difference ∆P mB<br />

Radius of maximum wind Rmax km<br />

2.2.3 Tropical Cyclone Track<br />

a = 4.5<br />

b = 6.75<br />

c = 9<br />

a = 15<br />

b = 22.5<br />

c = 30<br />

a = 20<br />

b = 35<br />

c = 64<br />

a = 4.5<br />

b = 6.75<br />

c = 9<br />

a = 45<br />

b = 57.5<br />

c = 60<br />

a = 13<br />

b = 32<br />

c = 64<br />

a = 4.5<br />

b = 6.75<br />

c = 9<br />

a = 85<br />

b = 95<br />

c = 105<br />

a = 8<br />

b = 20<br />

c = 40<br />

Two simplifying assumptions are made in modeling the tropical cyclone track. First, <strong>for</strong><br />

the purpose of estimating the wind damage to an individual structure, it is unnecessary<br />

to track the tropical cyclone from genesis to dissipation. We are only interested in the<br />

time period when the wind speed is above a certain threshold (below which damage is<br />

highly unlikely). It is assumed that the tropical cyclone track remains straight during this<br />

relatively short time period. Second, we are interested in the wind angle relative to the<br />

structure’s orientation, not the absolute wind angle with respect to the cardinal directions;<br />

there<strong>for</strong>e the absolute translational direction of the tropical cyclone does not matter.<br />

Based on these assumptions, the location of the tropical cyclone track can be fully<br />

defined by the closest distance between the track <strong>and</strong> the structure, denoted herein by s. A<br />

plus or minus sign is assigned to the quantity s to indicate whether the structure is on the


Chapter 2. Structural Components <strong>and</strong> <strong>Wind</strong> Directionality 19<br />

right or the left side with respect to the translational direction of the tropical cyclone. The<br />

condition s = 0 means that the tropical cyclone center passes directly over the structure.<br />

Let t = 0 indicate the time when the tropical cyclone track is closest to structure. As<br />

illustrated in Figure 2.2, at time t the distance between the storm center <strong>and</strong> the structure<br />

location is:<br />

d(t) =<br />

<br />

V 2<br />

t t2 + s 2 , (2.1)<br />

<strong>and</strong> the relative angle from the structure location to the storm center at time t is:<br />

2.2.4 Mean Boundary Layer <strong>Wind</strong><br />

−1 s<br />

φ(t) = tan . (2.2)<br />

Vtt<br />

Given d(t) <strong>and</strong> φ(t), the wind velocity at the location of the structure can be characterized<br />

by inserting the sampled tropical cyclone properties into a wind field model. Instead of<br />

solving the wind field by differential equations based on Navier-Stokes, we calculate the<br />

wind velocity over time using the framework developed by Jakobsen <strong>and</strong> Madsen (2004).<br />

In a tropical cyclone event, the mean boundary layer wind is composed of the environ-<br />

mental-scale wind <strong>and</strong> the storm-scale wind. The storm-scale wind can be further broken<br />

down into tangential <strong>and</strong> radial wind speed components. This can be expressed by<br />

Vmbd = Venv + Vtan + Vrad, (2.3)<br />

where the subscript ‘mbd’ abbreviates ‘mean boundary layer’. The <strong>for</strong>mulation of each<br />

component, according to Jakobsen <strong>and</strong> Madsen (2004), is briefly described below.<br />

The environmental-scale wind, Venv, at the location of the structure has the same direc-<br />

tion as the translational direction of the storm (see Figure 2.2). Its magnitude is expressed


as:<br />

Chapter 2. Structural Components <strong>and</strong> <strong>Wind</strong> Directionality 20<br />

<br />

Venv(d) = Vt · exp − d<br />

<br />

, (2.4)<br />

RG<br />

where RG, the length scale of the environmental-scale processes, is taken to be 500 km<br />

(Jakobsen <strong>and</strong> Madsen, 2004). The tangential wind speed, Vtan, is calculated based on<br />

the parametric tropical cyclone model developed by Jakobsen <strong>and</strong> Madsen (2004) <strong>and</strong> the<br />

air pressure distribution model described by Holl<strong>and</strong> (1980). The latter expresses the air<br />

pressure as follows:<br />

P (r) = Pc + ∆P exp<br />

<br />

−<br />

Rmax<br />

r<br />

B <br />

, (2.5)<br />

where r is distance from the storm center, Pc is the central pressure, <strong>and</strong> B is a dimension-<br />

less scaling parameter (affecting the shape of the pressure profile). A simple empirical<br />

equation is used to relate B to Pc, expressed in mbar (Hubbert et al., 1991):<br />

B = 1.5 +<br />

980 − Pc<br />

. (2.6)<br />

120<br />

This expression, originally derived based on data from Australian tropical cyclones but<br />

found to be generally applicable, is adopted herein mainly <strong>for</strong> its simplicity. The magnitude<br />

of the tangential wind at the structure’s location is (Jakobsen <strong>and</strong> Madsen, 2004):<br />

Vtan(d) =<br />

λd<br />

P<br />

· ∂P<br />

∂r<br />

+ 1<br />

4 f 2d2 − 1<br />

2<br />

fd, (2.7)<br />

where β is a parameter relating tangential <strong>and</strong> radial wind speeds <strong>and</strong> f is the Coriolis<br />

parameter. In the Northern hemisphere, the tangential wind is counter-clockwise, so the<br />

tangential wind angle at the structure location is φ − π/2. The radial wind, Vrad, at the<br />

structure’s location has a wind angle equal to φ + π. Its magnitude is based on that of the


Chapter 2. Structural Components <strong>and</strong> <strong>Wind</strong> Directionality 21<br />

tangential wind speed (Jakobsen <strong>and</strong> Madsen, 2004):<br />

Vrad(d) = Vtan(d) · B(Bd′−2B + (1 − 3B)d ′−B + B − 1)k − 2k − 2d ′ c<br />

B(d ′−B − 1) + 2 + 2fdV −1<br />

tan (d)<br />

, (2.8)<br />

where d ′ = d/Rmax, k is a diffusion parameter, <strong>and</strong> c is a drag parameter.<br />

V (m/s)<br />

(deg)<br />

60<br />

40<br />

20<br />

0<br />

180<br />

90<br />

0<br />

90<br />

5 10 15 20<br />

t (hr)<br />

180<br />

0 5 10 15 20<br />

t (hr)<br />

V (m/s)<br />

(deg)<br />

60<br />

40<br />

20<br />

0<br />

180<br />

90<br />

0<br />

90<br />

5 10 15<br />

t (hr)<br />

180<br />

0 5 10 15<br />

t (hr)<br />

(a) (b)<br />

Figure 2.3: Sample simulated wind speed time series (V (t)) <strong>and</strong> wind angle time series<br />

(θ(t)) using parameters: (a) Vt = 7 m/s, ∆P = 55 mB, Rmax = 50 km, <strong>and</strong> s = -50 km;<br />

(b) Vt = 12 m/s, ∆P = 47 mB, Rmax = 80 km, <strong>and</strong> s = 30 km.<br />

Figure 2.3 shows a sample of the wind (speed <strong>and</strong> orientation) profiles at two loca-<br />

tions during a tropical cyclone. Parameters such as the Coriolis <strong>for</strong>ce, diffusion coefficient,<br />

frictional drag coefficient are assumed to be constant <strong>for</strong> each sampled tropical cyclone.<br />

2.2.5 Three-Second Surface <strong>Wind</strong><br />

The mean boundary wind velocity Vmbd at the location of the structure needs to be converted<br />

to the near-surface (10-meter) wind velocity <strong>for</strong> wind-load calculation. First, the 0.8 mean<br />

conversion factor is applied to obtain the best estimate of the surface wind velocity over<br />

open water (Powell, 1980):<br />

Vwater = 0.8 Vmbd. (2.9)


Chapter 2. Structural Components <strong>and</strong> <strong>Wind</strong> Directionality 22<br />

It is assumed that represents the mean wind speed over a 1-hour time period (Vickery<br />

<strong>and</strong> Twisdale, 1995); it is denoted by , where 3600 is the averaging time in seconds. The<br />

quantity V water<br />

3600 can be further converted to a best estimate of the three-second wind speed<br />

at 10 m over open-terrain exposure (Simiu et al., 2007):<br />

V l<strong>and</strong><br />

3<br />

= (1.16)(1.07) V water<br />

3600 . (2.10)<br />

The three-second wind speed is defined as the highest average speed, when averaging is over<br />

segments of three-second duration. Note that V l<strong>and</strong><br />

3<br />

is the wind speed used in wind load<br />

calculation according to the ASCE 7 St<strong>and</strong>ard (American Society of Civil Engineers, 2003).<br />

Combining all the factors yields a direct relationship between V l<strong>and</strong><br />

3<br />

<strong>and</strong> Vmbd :<br />

V l<strong>and</strong><br />

3 = 0.993 Vmbd. (2.11)<br />

This simple approximate expression does of course not consider any specific topographic<br />

features of the area of interest. It assumes that the structure is located in terrain with open<br />

exposure. For the sake of clarity, V l<strong>and</strong><br />

3<br />

2.2.6 Damage Threshold<br />

will be denoted V in what follows.<br />

As stated earlier, our interest focuses on the parts of wind records when the wind speed<br />

exceeds a certain damage-related threshold. The level selected is 20 m/s, based on the lowest<br />

wind speed (50 mph) needing to be considered by the Florida Department of Financial<br />

Services in their hurricane loss model (Gurley et al., 2006). Hence, when a tropical cyclone<br />

is approaching the structure, the wind speed <strong>and</strong> wind angle begin to be re-corded when<br />

the wind speed exceeds the damage threshold, <strong>and</strong> the record stops when the cyclone is far<br />

enough away so that no subsequent wind speed exceeds 20 m/s. In each time series, the<br />

time when the wind speed first reaches 20 m/s is set at zero.


Chapter 2. Structural Components <strong>and</strong> <strong>Wind</strong> Directionality 23<br />

In situations when the distance s is large <strong>and</strong> the intensity of the tropical cyclone low,<br />

it may happen that the wind speed never exceeds the damage threshold. In this case, no<br />

wind speed or wind angle is recorded. The simulation record is then labeled ‘below damage<br />

threshold’; the wind loads in these records are assumed to be negligible.<br />

V (m/s)<br />

(deg)<br />

60<br />

40<br />

20<br />

0<br />

180<br />

90<br />

0<br />

90<br />

5 10 15 20<br />

t (hr)<br />

180<br />

0 5 10 15 20<br />

t (hr)<br />

V (m/s)<br />

(deg)<br />

0<br />

180<br />

(a) (b)<br />

60<br />

40<br />

20<br />

90<br />

0<br />

90<br />

5 10 15<br />

t (hr)<br />

180<br />

0 5 10 15<br />

t (hr)<br />

Figure 2.4: <strong>Wind</strong> speed <strong>and</strong> wind angle time records during Hurricane Katrina <strong>and</strong> Wilma:<br />

(a) Data record at Belle Chasse, LA during Hurricane Katrina; (b) Data record at Everglade<br />

City, FL during Hurricane Wilma (Civil <strong>and</strong> Coastal Engineering Department of the<br />

University of Florida, 2009).<br />

2.2.7 Sample Results<br />

Two sample sets of time series generated by the probabilistic wind model described above are<br />

compared with some historical hurricane wind (speed <strong>and</strong> orientation) records of the Florida<br />

Coastal Monitoring Program (FCMP) at the University of Florida (Civil <strong>and</strong> Coastal En-<br />

gineering Department of the University of Florida, 2009). The three hurricane parameters<br />

(Vt, ∆P <strong>and</strong> Rmax) used to generate the two sample sets are obtained from the tropical cy-<br />

clone reports of the National Hurricane Center (Knabb et al. 2006; Pasch et al. 2006). The<br />

relative distance between the tropical cyclone track <strong>and</strong> the record location (s) is obtained<br />

from the FCMP tower deployment map (Civil <strong>and</strong> Coastal Engineering Department of the


Chapter 2. Structural Components <strong>and</strong> <strong>Wind</strong> Directionality 24<br />

University of Florida, 2009). The values of the parameters are listed in the caption of Figure<br />

2.3, which shows the simulated wind records. For comparison, the historical wind records<br />

(see Figure 2.4) are converted from 15-minute average to 3-second gust. The general trends<br />

of the simulated wind speed <strong>and</strong> wind angle time series match the historical record quite<br />

well. The magnitude of the wind velocity is also close to the historical data. Note that the<br />

absolute accuracy of this model is not the major focus in this study. Since the objective is<br />

to look at the effect of changing wind directionality in individual wind events, it is more<br />

important <strong>for</strong> the wind time series model to be capable to generate wind time series that<br />

have trends similar to those of the historical wind records, in terms of both wind speed <strong>and</strong><br />

wind angle.<br />

2.3 Alternative Approaches to <strong>Wind</strong>-Load <strong>and</strong> Reliability<br />

<strong>Assessment</strong><br />

Given the simulated tropical cyclone wind velocity time series described in the previous<br />

section, wind loads may be computed <strong>and</strong> structural reliability analysis may be carried out,<br />

<strong>for</strong> the two a<strong>for</strong>e-mentioned approaches, in the way explained below.<br />

2.3.1 <strong>Wind</strong> Load <strong>and</strong> Reliability Calculation<br />

Damage of envelope components, which is the major factor in insurance claims (Sparks<br />

et al., 1994), is mainly caused by wind loads during tropical cyclones (National Institute of<br />

St<strong>and</strong>ards <strong>and</strong> Technology, 2006b). Given a wind velocity time series over the time domain<br />

[0, tL] (with wind speed V (t) <strong>and</strong> θ(t)), from bluff-body aerodynamics, the wind pressure<br />

acting on a building envelope component at time t ∈ [0, tL] is given by (Holmes, 2007):<br />

S(t) = 1<br />

2 ρV (t)2 Cp(θ(t) − γ), (2.12)


Chapter 2. Structural Components <strong>and</strong> <strong>Wind</strong> Directionality 25<br />

where ρ is the air density, <strong>and</strong> Cp is the non-dimensional pressure coefficient of the compo-<br />

nent. The pressure coefficient, which depends on the size <strong>and</strong> orientation of the structure<br />

<strong>and</strong> the structural component, is derived from wind tunnel tests. It is generally modeled as<br />

a function of θ(t) with respect to the component orientation, γ.<br />

Assume that wind pressure is the only damage source to the structural component. The<br />

limit state may be defined such that S is the wind pressure acting on the component <strong>and</strong> R<br />

is the capacity to resist the wind pressure. It is assumed that dead load is relatively small<br />

<strong>and</strong> may be (conservatively) ignored. Since tL is very short compared to the lifetime of the<br />

structure, deterioration is assumed to be negligible across [0, tL]. Thus, R may be defined<br />

as a r<strong>and</strong>om variable whose value remains constant over time. The probability of failure<br />

may be written as<br />

Pf = P {S(t) − R > 0, <strong>for</strong> any t ∈ [0, tL]}. (2.13)<br />

Since R is constant over time, the probability of failure may also be expressed as:<br />

where Smax is the maximum value of S within [0, tL].<br />

Pf = P {Smax − R}, (2.14)<br />

2.3.2 Time-Stepping Method <strong>and</strong> Point-in-Time Method<br />

In the time-stepping method, a sample wind load S(t) is evaluated at time steps at which<br />

the wind time series sample values of V (t) <strong>and</strong> θ(t) are recorded. It is then compared with<br />

R at each of those time steps. If S(t) is larger than R at any time step, then the building<br />

component fails <strong>for</strong> that specific wind time series sample (equation (2.13)). Alternatively,<br />

<strong>for</strong> each sample wind time series, Smax may be found by comparing the values of S(t) at<br />

every time step, <strong>and</strong> then compared to R (equation (2.14)).<br />

The point-in-time method re<strong>for</strong>mulates the problem by evaluating the probability of


Chapter 2. Structural Components <strong>and</strong> <strong>Wind</strong> Directionality 26<br />

failure at a single point in time, t ∗ , within the time domain [0, tL]. (This approach is in a<br />

sense compatible with the characteristics of wind loads on low-rise structures, when dynamic<br />

response amplification is insignificant.) The method approximates the probability of failure<br />

as:<br />

Pf ≈ P ∗ f = Prob{S(t∗ ) − R < 0}. (2.15)<br />

The above equation would be equivalent to equation (2.13) if S(t ∗ ) = Smax. However,<br />

finding t ∗ such that S(t ∗ ) = Smax requires evaluation of S(t) across the time domain [0, tL],<br />

which is equivalent to the time-stepping method described above. To simplify the problem,<br />

the effect of wind direction is assumed to be negligible so that t ∗ is selected as the time<br />

point at which the wind speed is maximum, e.g. V (t ∗ ) = maxt V (t). There<strong>for</strong>e,<br />

2.4 Numerical Examples<br />

2.4.1 Setup<br />

Smax ≈ S ∗ = 1<br />

2 ρV (t∗ ) 2 Cp(θ(t ∗ ) − γ). (2.16)<br />

This section focuses on the building envelope of low-rise residences, which is the most vul-<br />

nerable building component in tropical cyclones under wind pressure (National Institute of<br />

St<strong>and</strong>ards <strong>and</strong> Technology, 2006b). Four pressure coefficients of building envelope compo-<br />

nents are selected to illustrate the difference in results between the two analysis methods;<br />

these coefficients relate to:<br />

1. Positive Wall Pressure;<br />

2. Negative Wall Pressure;<br />

3. External Pressure at Roof Corner;<br />

4. External Pressure at Near-Edge Roof.


Chapter 2. Structural Components <strong>and</strong> <strong>Wind</strong> Directionality 27<br />

Pressure Coefficient<br />

4<br />

3<br />

2<br />

1<br />

Negative Wall Pressure<br />

Positive Wall Pressure<br />

External Pressure<br />

at Roof Corner<br />

External<br />

Pressure<br />

at Roof<br />

NearEdge<br />

0<br />

0 90 180<br />

<strong>Wind</strong> Angle (deg)<br />

270 360<br />

Figure 2.5: Four selected pressure coefficients with respect to wind angle.<br />

The pressure coefficients (see Figure 2.5) are all obtained from the HAZUS-MH3 tech-<br />

nical manual (Federal Emergency <strong>Management</strong> Agency, 2006) <strong>and</strong> the Florida Public Hur-<br />

ricane Loss Projection Model engineering team report (Gurley et al., 2006). This study<br />

focuses on the basic cases of evaluating wind pressure using pressure coefficient functions.<br />

Reliability analysis of structural systems is not explicitly considered in this study, as this<br />

would involve many more factors, such as load paths, pressure zone distribution, <strong>and</strong> de-<br />

pendence among component failures, thereby clouding the focus on comparing the time-<br />

stepping <strong>and</strong> point-in-time methods <strong>and</strong> quantifying the effect of changing wind direction-<br />

ality. Comparison, as regards structural system reliability, between the time-stepping <strong>and</strong><br />

point-in-time methods is dealt with in the work by Lin et al. (2010) <strong>and</strong> Yau et al. (in<br />

press).<br />

2.4.2 Maximum <strong>Wind</strong> Pressure<br />

10,000 wind velocity time series were generated at one-minute intervals <strong>for</strong> each of the<br />

hurricane categories 1, 3 <strong>and</strong> 5 by r<strong>and</strong>omly sampling Vt, ∆P <strong>and</strong> Rmax using the probability


Chapter 2. Structural Components <strong>and</strong> <strong>Wind</strong> Directionality 28<br />

distribution parameters by Iman et al. (2002) <strong>and</strong> assuming that s is uni<strong>for</strong>mly distributed<br />

from -200 km to 200 km. This results in three collections of simulated wind time series<br />

in which the tropical cyclone track is at most 200 km away from the structure. Note<br />

that since the entire analytical model involves only algebra, a large number of simulated<br />

wind time series can be generated with scant computational ef<strong>for</strong>t. The three wind time<br />

series databases are then employed to calculate the wind pressure over time associated with<br />

each pressure coefficient. Each database, generated with the category 1, 3 or 5 hurricane<br />

parameters, consists of 10,000 wind time series <strong>and</strong> represents the scenario in which the<br />

tropical cyclone track is at most 200 km from the structure. The angle in which the<br />

tropical cyclone approaches the structure is assumed to be uni<strong>for</strong>mly distributed within<br />

[0, 2π]. For each wind time series, we fix the absolute wind direction <strong>and</strong> let the orientation<br />

of the structure be one of the 24 evenly spaced angles across [0, 2π]. As a result, 240,000<br />

individual wind velocity time series are considered <strong>for</strong> each hurricane category.<br />

Figure 2.6 shows the cumulative distribution functions of maximum wind pressure over<br />

time using the hurricane-category-5 simulated wind time series database. In each plot, the<br />

solid line is the cumulative distribution of the maximum wind pressure over time (Smax)<br />

obtained by the time-stepping method, <strong>and</strong> the dashed line is the cumulative distribution<br />

of wind pressure at maximum wind speed (S ∗ ) obtained by the point-in-time approach.<br />

The mean values of both distributions are shown in the plots. The cumulative distributions<br />

using hurricane categories 1 <strong>and</strong> 3 are not shown due to space constraints.<br />

The difference between Smax <strong>and</strong> S ∗ , indicating the degree of inaccuracy of the point-<br />

in-time approach, varies with the pressure coefficient. To compare the two cumulative<br />

distributions of each wind pressure quantitatively, the Kolmogorov-Smirnov statistic is cal-<br />

culated. It is defined (Kendall et al., 1999) as the largest absolute difference between two


Chapter 2. Structural Components <strong>and</strong> <strong>Wind</strong> Directionality 29<br />

cumulative distribution functions, F1(x) <strong>and</strong> F2(x), at the same value of the variable:<br />

K-S Statistic = max<br />

−∞


NEGATIVE WALL PRESSURE<br />

(COV = 0.509)<br />

POSITIVE WALL PRESSURE<br />

(COV = 0.977)<br />

Chapter 2. Structural Components <strong>and</strong> <strong>Wind</strong> Directionality 30<br />

0.5<br />

0.5<br />

0.4<br />

0.4<br />

0.3<br />

0.2<br />

K-S Statistic<br />

0.3<br />

0.2<br />

K-S Statistic<br />

0.1<br />

0.1<br />

0<br />

All Records |s| – Rmax < 0 |s| – Rmax ! 0<br />

EXTERNAL PRESSURE AT ROOF NEAR-<br />

EDGE<br />

ROOF NEAR-EDGE EXTERNAL (COV PRESSURE = 0.093)<br />

0.5 (COV = 0.093)<br />

0<br />

All Records |s| – Rmax < 0 |s| – Rmax ! 0<br />

EXTERNAL PRESSURE AT ROOF CORNER<br />

(COV = 0.200)<br />

0.5<br />

0.5<br />

LEGEND<br />

0.4<br />

0.3<br />

0.4<br />

0.4<br />

Category 1<br />

Category 3<br />

0.2<br />

K-S Statistic<br />

0.3<br />

0.2<br />

K-S Statistic<br />

0.3<br />

Category 5<br />

0.1<br />

0.2<br />

0<br />

K-S Statistic<br />

|s| – Rmax !<br />

0<br />

All Records |s| – Rmax <<br />

0<br />

0.1<br />

0.1<br />

0<br />

0<br />

All Records |s| – Rmax < 0 |s| – Rmax ! 0<br />

All Records |s| – Rmax < 0 |s| – Rmax ! 0<br />

Figure 2.7: Kolmogorov-Smirnov statistics comparing the maximum wind load (Smax) (obtained from time-stepping<br />

method) <strong>and</strong> the approximated maximum wind load (S∗ ) (obtained from point-in-time method).


Chapter 2. Structural Components <strong>and</strong> <strong>Wind</strong> Directionality 31<br />

A detailed analysis is conducted to examine how the Kolmogorov-Smirnov statistic<br />

changes under different conditions. The results of the analysis are illustrated in Figure<br />

2.7. First, we are interested in how the pressure coefficient function affects the results. The<br />

variation of pressure coefficient with respect to wind angle is quantified by the coefficient of<br />

variation, which is shown in the title of each plot in Figure 2.7. The plots show a very high<br />

correlation between the Kolmogorov-Smirnov statistic <strong>and</strong> the coefficient of variation of the<br />

pressure coefficients, confirming that the variation in the pressure coefficient, indicative of<br />

the sensitivity of wind pressure to wind angle, is one of the main sources of inaccuracy of<br />

the point-in-time approach.<br />

Second, we are interested in the effect of tropical cyclone track location on the results.<br />

For here we categorize all the wind time series into two types, depending on whether the<br />

location of the structure is ever inside the radius of maximum wind Rmax of a tropical<br />

cyclone. This condition may be indicated by the difference between |s| <strong>and</strong> Rmax:<br />

1. |s| − Rmax < 0<br />

The tropical cyclone track is close enough to the structure, so the latter is within<br />

Rmax <strong>for</strong> a period of time. The wind speed series is likely to be M-shaped <strong>and</strong> the<br />

change in wind angle is abrupt, as shown in Figure 2.4b.<br />

2. |s| − Rmax ≥ 0<br />

The structure is never inside Rmax. The wind speed series is likely to be bell-shaped<br />

<strong>and</strong> the change in wind angle is more gradual, as shown in Figure 2.4a.<br />

Note that |s| − Rmax is not a precise parameter that unambiguously indicates whether<br />

a wind speed time series has a M-shape, owing to the asymmetry of the tropical cyclone<br />

structure (higher wind speed on the left side <strong>and</strong> lower wind speed on the right side in the<br />

Northern hemisphere). For example, if the structure is to the left of the cyclone path <strong>and</strong><br />

is inside Rmax, the trend of the wind speed time series may not show any M-shape.


Chapter 2. Structural Components <strong>and</strong> <strong>Wind</strong> Directionality 32<br />

From Figure 2.7, the Kolmogorov-Smirnov statistics are significantly larger in the case<br />

when |s| − Rmax < 0 <strong>for</strong> each of the four pressure coefficients, indicating that the point-in-<br />

time approach underestimates the wind pressure more if the structure is ever inside Rmax.<br />

This shows that the different trends of wind speed <strong>and</strong> wind angle due to the track location<br />

affects the accuracy of the point-in-time approach. A more abrupt change in wind speed<br />

<strong>and</strong> wind angle tends to yield less accurate results.<br />

Third, we examine the effect of hurricane intensity on the results. Considering all records<br />

(without categorizing the data by track location), Kolmogorov-Smirnov statistics de-crease<br />

as one goes from category 1 to category 5 <strong>for</strong> all four pressure coefficients. This indicates<br />

that the point-in-time approach is less accurate in calculating wind pressure <strong>for</strong> hurricanes<br />

of lower intensity. However, further breaking down the results by examining how the K-S<br />

statistics change <strong>for</strong> each case separately, |s| − Rmax < 0 <strong>and</strong> |s| − Rmax ≥ 0, one finds that<br />

the inaccuracy of the point-in-time approach grows with the tropical cyclone intensity <strong>for</strong><br />

|s| − Rmax < 0, <strong>and</strong> vice versa <strong>for</strong> |s| − Rmax ≥ 0. Since the |s| − Rmax ≥ 0 records dominate<br />

<strong>for</strong> all hurricane categories (see Table 2.2 <strong>for</strong> the ratio of data in each category), on the<br />

whole, the inaccuracy of the point-in-time approach decreases as tropical cyclone intensity<br />

increases.<br />

Table 2.2: Percentage of simulated wind records by track location.<br />

|s| − Rmax < 0 |s| − Rmax ≥ 0<br />

Category 1 36.90% 63.10%<br />

Category 3 26.31% 73.69%<br />

Category 5 18.29% 81.71%<br />

2.4.3 Probability of Failure<br />

The next step is to apply the calculated wind pressure to reliability analysis. Here we focus<br />

on single-limit-state failures. We computed two probabilities of failure <strong>for</strong> each of the four


Chapter 2. Structural Components <strong>and</strong> <strong>Wind</strong> Directionality 33<br />

types of wind pressure by assuming two different probability models <strong>for</strong> the resistances:<br />

normal <strong>and</strong> lognormal. While the normal distribution is widely adopted in current vulner-<br />

ability models, including the HAZUS-MH model (Vickery et al., 2006b) <strong>and</strong> the Florida<br />

Public Hurricane Loss Projection Model (Gurley et al., 2006), the lognormal distribution<br />

is generally considered a better model <strong>for</strong> resistance. For comparison, the mean <strong>and</strong> stan-<br />

dard deviation is set at the same values in both models. The probability of failure of a<br />

roof sheathing panel resisting the external pressure at a roof corner or near-edge roof is<br />

computed, by assuming that the resistance capacity of the panel has mean <strong>and</strong> st<strong>and</strong>ard<br />

deviation equal to 6,000 Pa <strong>and</strong> 2,400 Pa, respectively. Similarly, the probability of failure<br />

of a wall sheathing panel resisting wall pressures is calculated by assuming that the resis-<br />

tance capacity of the panel has mean <strong>and</strong> st<strong>and</strong>ard deviation equal to 7,000 Pa <strong>and</strong> 2,800<br />

Pa, respectively. From equation (2.14), the probability of failure can be expressed as:<br />

∞<br />

Pf = FR(x)fSmax(x)dx (2.18)<br />

−∞<br />

In this case, we have 240,000 r<strong>and</strong>om samples of Smax (<strong>and</strong> S ∗ ) <strong>and</strong> a complete probabilistic<br />

description of R, There<strong>for</strong>e, the probability of failure can be calculated from:<br />

Pf ≈<br />

n<br />

i=1<br />

FR(Smax,i)<br />

n<br />

(2.19)<br />

where n is the total number of samples of Smax (240,000), FR(·) is the cumulative distribu-<br />

tion of R, <strong>and</strong> Smax, i is the i-th r<strong>and</strong>om sample of Smax.<br />

For the case of normally distributed resistances, the results of the reliability analysis<br />

are shown in Table 2.3. The percentage difference in failure probabilities between the<br />

time-stepping <strong>and</strong> point-in-time methods is compared, as illustrated in Figure 2.8. For the<br />

case of lognormally distributed resistances, the results are shown indirectly, as percentage<br />

differences in failure probabilities relative to the normal-distribution case, in Figure 2.9.


Chapter 2. Structural Components <strong>and</strong> <strong>Wind</strong> Directionality 34<br />

Table 2.3: Probabilities of failure of four component limit states<br />

Time-Stepping Method Point-in-Time Method<br />

Category<br />

All records |s| − Rmax < 0 |s| − Rmax ≥ 0 All records |s| − Rmax < 0 |s| − Rmax ≥ 0<br />

1 16.88% 51.80% 9.051% 10.40% 28.052% 6.459%<br />

Positive<br />

3 6.93% 15.47% 8.951% 9.43% 25.260% 5.885%<br />

Wall<br />

5 1.65% 2.20% 9.406% 14.06% 40.268% 8.193%<br />

Pressure<br />

1 16.88% 51.80% 9.051% 10.40% 28.052% 6.459%<br />

Positive<br />

3 6.93% 15.47% 8.951% 9.43% 25.260% 5.885%<br />

Wall<br />

5 1.65% 2.20% 9.406% 14.06% 40.268% 8.193%<br />

Pressure<br />

1 16.88% 51.80% 9.051% 10.40% 28.052% 6.459%<br />

Positive<br />

3 6.93% 15.47% 8.951% 9.43% 25.260% 5.885%<br />

Wall<br />

5 1.65% 2.20% 9.406% 14.06% 40.268% 8.193%<br />

Pressure<br />

1 16.88% 51.80% 9.051% 10.40% 28.052% 6.459%<br />

Positive<br />

3 6.93% 15.47% 8.951% 9.43% 25.260% 5.885%<br />

Wall<br />

5 1.65% 2.20% 9.406% 14.06% 40.268% 8.193%<br />

Pressure


NEGATIVE WALL PRESSURE<br />

(COV = 0.509)<br />

POSITIVE WALL PRESSURE<br />

(COV = 0.977)<br />

Chapter 2. Structural Components <strong>and</strong> <strong>Wind</strong> Directionality 35<br />

-60%<br />

-60%<br />

-40%<br />

-40%<br />

-20%<br />

Percentage Difference<br />

-20%<br />

Percentage Difference<br />

0%<br />

All Records |s| – Rmax < 0 |s| – Rmax ! 0<br />

EXTERNAL PRESSURE AT ROOF NEAR-<br />

EDGE<br />

NEAR-EDGE ROOF EXTERNAL (COV PRESSURE = 0.093)<br />

0.5 (COV = 0.093)<br />

0%<br />

All Records |s| – Rmax < 0 |s| – Rmax ! 0<br />

ROOF CORNER EXTERNAL PRESSURE<br />

(COV = 0.200)<br />

LEGEND<br />

0.4<br />

-60%<br />

-60%<br />

0.3<br />

Category 1<br />

Category 3<br />

0.2<br />

K-S Statistic<br />

-40%<br />

-40%<br />

Category 5<br />

0.1<br />

0<br />

-20%<br />

-20%<br />

|s| – Rmax !<br />

0<br />

All Records |s| – Rmax <<br />

0<br />

Percentage Difference<br />

Percentage Difference<br />

0%<br />

0%<br />

All Records |s| – Rmax < 0 |s| – Rmax ! 0<br />

All Records |s| – Rmax < 0 |s| – Rmax ! 0<br />

Figure 2.8: Percentage differences between the probabilities of failure (Pf ) (obtained from time-stepping method) <strong>and</strong> the<br />

approximated probabilities of failure (P ∗ f ) obtained from point-in-time method) <strong>for</strong> normally distributed resistances.


NEGATIVE WALL PRESSURE<br />

(COV = 0.509)<br />

POSITIVE WALL PRESSURE<br />

(COV = 0.977)<br />

Chapter 2. Structural Components <strong>and</strong> <strong>Wind</strong> Directionality 36<br />

-60%<br />

-60%<br />

-40%<br />

-40%<br />

-20%<br />

Percentage Difference<br />

-20%<br />

Percentage Difference<br />

0%<br />

All Records |s| – Rmax < 0 |s| – Rmax ! 0<br />

EXTERNAL PRESSURE AT ROOF NEAR-<br />

EDGE<br />

NEAR-EDGE ROOF EXTERNAL (COV PRESSURE = 0.093)<br />

0.5 (COV = 0.093)<br />

0%<br />

All Records |s| – Rmax < 0 |s| – Rmax ! 0<br />

ROOF CORNER EXTERNAL PRESSURE<br />

(COV = 0.200)<br />

LEGEND<br />

0.4<br />

-60%<br />

-60%<br />

0.3<br />

Category 1<br />

Category 3<br />

0.2<br />

K-S Statistic<br />

-40%<br />

-40%<br />

Category 5<br />

0.1<br />

0<br />

-20%<br />

-20%<br />

|s| – Rmax !<br />

0<br />

All Records |s| – Rmax <<br />

0<br />

Percentage Difference<br />

Percentage Difference<br />

0%<br />

0%<br />

All Records |s| – Rmax < 0 |s| – Rmax ! 0<br />

All Records |s| – Rmax < 0 |s| – Rmax ! 0<br />

Figure 2.9: Percentage differences between the probabilities of failure (Pf ) (obtained from time-stepping method) <strong>and</strong> the<br />

approximated probabilities of failure (P ∗ f ) (obtained from point-in-time method) <strong>for</strong> lognormally distributed resistances.


Chapter 2. Structural Components <strong>and</strong> <strong>Wind</strong> Directionality 37<br />

Overall, all probabilities of failure are underestimated significantly by the point-in-time<br />

method due to the underestimation of wind load. The degree of underestimation ranges from<br />

about -5% to -60%. In addition, by comparing Figures 2.8 <strong>and</strong> 2.9, it shows that the choice<br />

of probability model <strong>for</strong> the resistance changes the percentage difference in probability of<br />

failure substantially. Unlike the Kolmogorov-Smirnov statistics related to wind pressure,<br />

the percentage differences in probability of failure do not exhibit any obvious trend with the<br />

variation of the pressure coefficient with respect to wind angle. The relationship between<br />

hurricane intensity <strong>and</strong> the percentage difference in probability of failure also differs <strong>for</strong> each<br />

wind pressure coefficient. However, it is observed that in almost all instances the percentage<br />

difference is larger in case the structure is within the radius of maximum wind <strong>for</strong> some time<br />

interval (|s| − Rmax < 0) than when the structure is located outside the radius of maximum<br />

wind (|s| − Rmax ≥ 0).<br />

2.5 Discussion<br />

The point-in-time method underestimates the wind load primarily because it neglects the<br />

effect of changing wind angle during individual wind time series. Maximum wind pressure<br />

usually does not occur at maximum wind speed if the pressure coefficient varies significantly<br />

with the wind angle or if the change in wind angle is abrupt over time. In the single-limit-<br />

state reliability problems, the probabilities of failure are underestimated by up to 60%. If<br />

the interdependence among failures of structural components is considered, the point-in-<br />

time approach is expected to become even less accurate. For example, in low-rise structures,<br />

the breakage of a window, which changes the internal pressure of the structure, would affect<br />

the subsequent pressure loading on the envelope components <strong>and</strong> hence significantly change<br />

their failure probabilities. The point-in-time approach cannot capture the subsequent effect<br />

of any component’s failure on the other components. In addition, the numerical examples<br />

consider wind pressure only. If debris impact is included, the effect of wind directionality


Chapter 2. Structural Components <strong>and</strong> <strong>Wind</strong> Directionality 38<br />

on the reliability results probably will be greater, since the chance of debris impact may<br />

strongly depend on wind direction. The time-stepping method is capable of considering<br />

the above effects by evaluating the structural loading in a series of time steps, but such<br />

evaluation can be very computationally intensive. By using approximate methods such as<br />

FORM or SORM instead of Monte Carlo simulation, the computation time <strong>for</strong> the time-<br />

stepping method may be shortened. However, assessing structural reliability under tropical<br />

cyclone conditions is highly complex <strong>and</strong> nonlinear <strong>and</strong> such approximate methods may<br />

provide inaccurate results.<br />

The number or pattern of time steps may be modified in the time-stepping method, per-<br />

haps to save computational ef<strong>for</strong>t. One way is to increase the length of all evenly spaced<br />

time time intervals. However, since maximum wind pressure <strong>and</strong> component failures are<br />

more likely to occur at higher wind speeds, optimality (in time stepping) could be pursued<br />

by making the time intervals shorter at the high wind speeds, <strong>and</strong> longer at lower wind<br />

speeds.<br />

An underlying assumption of the analysis presented is that no in<strong>for</strong>mation is available<br />

regarding the orientation of the structure or the direction of the tropical cyclone track.<br />

When considering a prototype structure, this assumption seems appropriate, <strong>and</strong> it is com-<br />

monly adopted in wind-related loss estimation <strong>for</strong> a large number of buildings (see Gurley<br />

et al., 2006). However, in light of the significant effect of wind directionality on wind loads<br />

<strong>and</strong> on structural reliability, demonstrated in this study, there clearly is merit to analyze<br />

the prototype building using the time-stepping method. To evaluate a large number of<br />

individual buildings with specific orientations <strong>and</strong> locations, computation-time issues may<br />

be more crucial <strong>and</strong> the point-in-time method may be more advantageous.<br />

The analysis of the effect of changing wind direction depends almost entirely on the<br />

pressure coefficient functions, which provide approximations <strong>for</strong> the actual wind pressures<br />

acting on a structure. However, different wind tunnel testing conditions often yield very


Chapter 2. Structural Components <strong>and</strong> <strong>Wind</strong> Directionality 39<br />

different results, implying significant uncertainty in pressure coefficient functions <strong>for</strong> the<br />

same structure under the same wind condition (Fritz et al., 2008; Ho et al., 2005; Pierre<br />

et al., 2005). Further work is needed to determine the impact of these errors (in pressure<br />

coefficient functions) on the effects of changing wind directionality.<br />

2.6 Closing Remarks<br />

This chapter presents a probabilistic model to generate tropical cyclone wind velocity time<br />

series <strong>for</strong> use in structural reliability analysis. With the databases of wind time series<br />

generated by this model, a detailed assessment is made of the differences between the results<br />

of the time-stepping <strong>and</strong> point-in-time methods. The study shows that the point-in-time<br />

approach significantly underestimates the wind load <strong>and</strong> the probability of failure. The<br />

degree of underestimation of maximum wind load is more significant when the variation<br />

of the pressure coefficient with respect to the wind angle is large, or when the structure<br />

is located inside the radius of maximum wind of a tropical cyclone <strong>for</strong> some period of<br />

time. As the tropical cyclone intensity increases, the degree of underestimation of the wind<br />

load generally decreases. When calculating component failure probabilities, the results are<br />

found to depend significantly on the choice of probabilistic model (normal vs. lognormal)<br />

<strong>for</strong> the resistance. Also, the accuracy of the point-in-time method is reduced in case the<br />

structure is located, <strong>for</strong> some period of time, within the radius of maximum wind of a<br />

tropical cyclone. Although the point-in-time method significantly underestimates both the<br />

wind load <strong>and</strong> the failure probability, it is more computationally efficient than the time-<br />

stepping approach. Future studies might pursue optimization of time-stepping procedures<br />

to best balance accuracy <strong>and</strong> computational efficiency.


Chapter<br />

3<br />

Internal Pressure in Low-Rise<br />

<strong>Structures</strong><br />

3.1 Introduction<br />

Design codes <strong>and</strong> st<strong>and</strong>ards, which are based on the results of various wind tunnel ex-<br />

periments <strong>and</strong> related research, provide a basis <strong>for</strong> the wind-related risk assessment of<br />

structures. However, current codes <strong>and</strong> st<strong>and</strong>ards do not completely cover all the factors<br />

involved in the complex structural damage process during high winds, such as the r<strong>and</strong>om-<br />

ness of wind pressure loads, quantification of windborne debris impact, progressive damage<br />

of different structural components, etc. More importantly, the purpose of the design codes<br />

<strong>and</strong> st<strong>and</strong>ards is to provide guidelines to avoid structural failure. Structural per<strong>for</strong>mance<br />

during <strong>and</strong> after failure is not the focus. To better assess the wind-related risk, structural<br />

models are developed to simulate the load paths <strong>and</strong> the failure sequence of some prototype<br />

low-rise residences (e.g. Unanwa et al., 2000; Gurley et al., 2005b). However, they require<br />

more available experimental data to be fully verified (Ellingwood et al., 2004).<br />

Besides the load paths inside the structural system <strong>and</strong> the building envelope system,<br />

internal pressure is the determining factor in the interplay between building components’<br />

40


Chapter 3. Internal Pressure in Low-Rise <strong>Structures</strong> 41<br />

failures. For example, during major wind events, the presence of a dominant opening<br />

on a building envelope may result in an increase in internal pressure, which significantly<br />

intensitifes the wind load on the roof (Holmes, 2007). Failures of the roof can further<br />

alter the internal pressure <strong>and</strong> cause more damages in other parts of the building. Past<br />

research ef<strong>for</strong>ts have indicated that internal pressure depends on several factors, including<br />

geometries of openings, overall building leakage, internal volume of the building, flexibility of<br />

the building envelope <strong>and</strong> structure, compartmentalization within the building, etc. (Ginger<br />

et al., 1997; Ginger, 2000; Kopp et al., 2008). Oh et al. (2007) summarized the recent work<br />

on internal pressure in houses.<br />

This chapter examines <strong>and</strong> discusses the relationship between the damage on building<br />

envelope <strong>and</strong> the internal pressure in low-rise structures. The scope of this study is confined<br />

to building envelope due to two reasons. First, as indicated by past hurricane damage sur-<br />

veys, hurricane winds primarily cause damage of components <strong>and</strong> cladding on the building<br />

envelope, e.g. roofing materials, windows, doors <strong>and</strong> garage doors (National Institute of<br />

St<strong>and</strong>ards <strong>and</strong> Technology, 2006b). The damage of envelope components, which causes rain<br />

to enter the building, is the major factor of insurance claims (Sparks et al., 1994). Second,<br />

the wind vulnerability of building envelope is studied extensively in the past literature. For<br />

example, Ellingwood et al. (2004) illustrated the sensitivity of roof panel <strong>and</strong> roof-to-wall<br />

connection fragility to exposure factor <strong>and</strong> component resistance. Lee <strong>and</strong> Rosowsky (2005)<br />

presented the reliability of roof panels given different roof shapes <strong>and</strong> structural geometry.<br />

Focusing on the vulnerability of building envelope is a continuation of past studies <strong>and</strong><br />

facilitates comparison with previous results.<br />

In this study, a structural model is developed to probabilistically simulate the wind load<br />

acting on roof panels <strong>and</strong> wall openings according to the modified ASCE 7 st<strong>and</strong>ard. Several<br />

methods of internal pressure adjustment are implemented into the model to <strong>for</strong>mulate a<br />

progressive damage mechanism, in which internal pressure can cause damage of structural


Chapter 3. Internal Pressure in Low-Rise <strong>Structures</strong> 42<br />

Figure 3.1: Unfolded view of baseline structures: (a) House A (b) House B (c) House C.


Chapter 3. Internal Pressure in Low-Rise <strong>Structures</strong> 43<br />

components, <strong>and</strong> any component failure may change back the internal pressure, causing<br />

further damages. The mathematical <strong>for</strong>mulation of the damage mechanism is presented. In<br />

the end, the effects of the different internal pressure adjustment methods on the structural<br />

vulnerability are closely examined <strong>and</strong> discussed. The significance of internal pressure in<br />

different conditions (presence of windborne debris impact, different number of windows,<br />

etc.) is also discussed.<br />

3.2 Structural Model<br />

Three baseline structures (see Figure 3.1) are considered in this study <strong>and</strong> they are based<br />

on a model that has been extensively studied in the <strong>Wind</strong> Load Test Facility at Clemson<br />

University (Rosowsky <strong>and</strong> Cheng, 1999a,b). Their roofs have been analyzed under a series<br />

of numerical fragility assessments (Lee <strong>and</strong> Rosowsky, 2005; Rosowsky <strong>and</strong> Cheng, 1999a,b).<br />

Different from the previous studies, we explicitly model the dimensions <strong>and</strong> locations of the<br />

wall openings including doors, windows, <strong>and</strong> garage doors. All structures are assumed to<br />

be located at an open-country terrain (exposure type C in ASCE 7 St<strong>and</strong>ard) (American<br />

Society of Civil Engineers, 2003).<br />

To access the vulnerability of the building envelope, three limit states are defined: roof<br />

sheathing uplift, opening pressure failure <strong>and</strong> opening debris impact failure. Each limit<br />

state function may be stated as R − (W + D), where R = resistance; W = wind load; D =<br />

dead load. The dead load is normally distributed with mean = 3.5 psf <strong>and</strong> COV = 0.1 <strong>for</strong><br />

roof sheathing (Lee <strong>and</strong> Rosowsky, 2005) <strong>and</strong> is equal to 0 <strong>for</strong> opening. The component is<br />

considered failed if the value of the limit state function is below zero.<br />

The resistance statistics of roof sheathing <strong>and</strong> openings are obtained from a previous<br />

study (Rosowsky <strong>and</strong> Cheng, 1999b) <strong>and</strong> the Florida Public Hurricane Loss Projection<br />

(FPHLP) Model (Gurley et al., 2005b), respectively. According to the studies, all the<br />

resistances are assumed normally distributed. A summary of resistance statistics is provided


in Table 3.1.<br />

Chapter 3. Internal Pressure in Low-Rise <strong>Structures</strong> 44<br />

Table 3.1: Statistics of structural component resistance.<br />

Component Dimension Limit State Unit Mean COV<br />

Roof sheathing panel 4’ × 8’ Uplift psf 57.7 0.20<br />

Roof sheathing panel 4’ × 4’ Uplift psf 73.3 0.20<br />

Glass window 3’6” × 3’6” Pressure psf 104.4 0.20<br />

Door 6’10” × 3’2” Pressure psf 100 0.20<br />

Garage door 6’10” × 12’ Pressure psf 52 0.20<br />

3.3 External <strong>Wind</strong> Pressure<br />

In ASCE 7 St<strong>and</strong>ard (American Society of Civil Engineers, 2003), the design wind pressure<br />

acting on components <strong>and</strong> cladding of a low-rise structure is determined from:<br />

W = qh(GCp − GCpi), (3.1)<br />

where qh = velocity pressure evaluated at mean roof height (h), G = gust factor; Cp =<br />

external pressure coefficient; Cpi = internal pressure coefficient. The velocity pressure<br />

evaluated at h is given by:<br />

qh = 0.00256KhKztKdV 2 I (units: lb/ft 2 ; V in mph), (3.2)<br />

where 0.00256 is half of the value of air density in English unit; Kh = velocity pressure<br />

exposure factor; Kzt = topographic factor; Kd = wind directionality factor; V = 3-second<br />

gust wind speed at 33 ft (10 m) <strong>and</strong> in open terrain; <strong>and</strong> I = importance factor. In<br />

our model, I is removed from equation (3.2) because it is a safety factor measuring the<br />

importance level of the structure. By assuming that there is no hill or escarpment around<br />

the structure, Kzt is also taken as unity. From a Delphi study by Ellingwood <strong>and</strong> Tekie


Chapter 3. Internal Pressure in Low-Rise <strong>Structures</strong> 45<br />

(1999), the velocity pressure exposure factor (Kh) is normally distributed (mean = 0.82,<br />

COV = 0.14) <strong>for</strong> exposure type C, <strong>and</strong> the directionality factor (Kd) is a normal r<strong>and</strong>om<br />

variable (mean = 0.89, COV = 0.16).<br />

External wind pressure varies spatially over the building envelope depending on the<br />

structural geometry <strong>and</strong> wind direction. Such spatial variation is represented by multiple<br />

pressure zones of different external pressure coefficient values (GCp) in ASCE 7 St<strong>and</strong>ard.<br />

Table 3.2 lists the statistics of pressure coefficient (GCp) of each pressure zone, which are<br />

obtained from the study by Lee <strong>and</strong> Rosowsky (2005), the FPHLP model Gurley et al.<br />

(2005b), <strong>and</strong> the study by Ellingwood <strong>and</strong> Tekie (1999). Figure 3.2 shows the distribution<br />

of pressure zones under two cases of wind directionality: (1) wind perpendicular to ridgeline<br />

<strong>and</strong> (2) wind parallel to ridgeline, according to Lee <strong>and</strong> Rosowsky (2005). By definition,<br />

the directionality factor (Kd) does not apply to both cases.<br />

External pressure on individual sheathing panel or wall opening is calculated by a<br />

weighted-average method. The value of external pressure coefficient on a component is<br />

the sum of all the external pressure coefficient on different pressure zones multiplied by the<br />

percentage of component area over which those pressures act on. The external pressure<br />

coefficient (GCp) values are r<strong>and</strong>omized <strong>for</strong> each component, according to the statistics in<br />

Table 3.2. In reality, the external pressure acting on the building envelope may change if<br />

the roof panels are uplifted. However, there is insufficient research data on such changes in<br />

external pressure. Thus, similar to many previous studies (e.g. Lee <strong>and</strong> Rosowsky, 2005),<br />

we assume that the external pressure is the same be<strong>for</strong>e <strong>and</strong> after the roof <strong>and</strong> the openings<br />

are damaged.<br />

3.4 Internal Pressure<br />

External pressure is often assumed constant regardless of the damage of the structure.<br />

However, internal pressure is often based upon the damage of building envelope in the


Chapter 3. Internal Pressure in Low-Rise <strong>Structures</strong> 46<br />

Figure 3.2: Component <strong>and</strong> cladding external pressure zones: (a) wind perpendicular to<br />

ridgeline (b) wind parallel to ridgeline. The wind directions are indicated by the arrows.<br />

The description of the pressure zones may be referred to Table 3.2.<br />

Table 3.2: Statistics of external pressure coefficient (GCp).<br />

Zone Location Nominal Mean COV<br />

1 Roof -0.9 -0.855 0.12<br />

2 Roof -1.7 -1.615 0.12<br />

3 Roof -2.6 -2.470 0.12<br />

4A <strong>Wind</strong>ward wall 1.0 0.950 0.12<br />

4B Leeward wall -0.8 -0.760 0.12<br />

4C Sidw wall -1.1 -1.045 0.12<br />

5 Side wall leading edge -1.4 -1.330 0.12


design guidelines.<br />

Chapter 3. Internal Pressure in Low-Rise <strong>Structures</strong> 47<br />

First, to calculate the initial internal pressure value of an intact structure, we employ the<br />

ASCE 7 St<strong>and</strong>ard (American Society of Civil Engineers, 2003). According to the St<strong>and</strong>ard,<br />

an intact structure is considered as an ‘enclosed building’, of which the internal pressure<br />

coefficient (GCpi) is ±0.18. Applying the statistics from Ellingwood <strong>and</strong> Tekie (1999), GCpi<br />

is assumed normally distributed (mean = ±0.15, COV = 0.33). Plus or minus value of the<br />

internal pressure, whichever yields a larger net pressure, is used to evaluate the failure status<br />

of a component.<br />

Once any building component fails, we need to adjust the internal pressure value ac-<br />

cording to the damage condition of the building envelope. Although the subject of internal<br />

pressure has been extensively studied, most studies consider only a single dominant wall<br />

opening (e.g. Ginger, 2000; Kopp et al., 2008), which does not fully represent many dam-<br />

age conditions in reality. There<strong>for</strong>e, in the following, we consider several internal pressure<br />

adjustment mechanisms adopted from different design codes <strong>and</strong> loss models, which are<br />

applicable to any envelope damage condition.<br />

3.4.1 Florida Public Hurricane Loss Projection (FPHLP) Model<br />

The FPHLP Model (Gurley et al., 2005b) models the internal pressure as a weighted average<br />

of external wind pressures acting on the failed wall openings. Each failed window <strong>and</strong> door<br />

is given an equal weight, while a failed garage door is given a weight four times the others<br />

due to its large surface area. The adjusted internal pressure coefficient is given by:<br />

<br />

GCpi =<br />

k wkGCp,k<br />

<br />

k wk<br />

, (3.3)<br />

where GCp,k is the external wind pressure coefficient acting on the k-th failed opening <strong>and</strong><br />

wk is the weight of the k-th failed opening. The weight wk is 4 if the k-th failed opening is<br />

the garage door, <strong>and</strong> is equal to 1 otherwise. It is assumed that a roof panel failure does


not affect the internal pressure.<br />

3.4.2 ASCE 7 St<strong>and</strong>ard<br />

Chapter 3. Internal Pressure in Low-Rise <strong>Structures</strong> 48<br />

In ASCE 7 st<strong>and</strong>ard (American Society of Civil Engineers, 2003), the value of internal<br />

pressure depends on the enclosure classification of a structure. A building is classified as a<br />

‘partially enclosed building’ if both the following conditions are satisfied (American Society<br />

of Civil Engineers, 2003):<br />

1. The total area of openings in a wall that receives positive external pressure exceeds<br />

the sum of the areas of openings in the balance of the building envelope (walls <strong>and</strong><br />

roof) by more than 10 percent.<br />

2. The total area of openings in a wall that receives positive external pressure exceeds<br />

4 ft 2 (0.37 m 2 ) or 1 percent of the area of that wall, whichever is smaller, <strong>and</strong> the<br />

percentage of openings in the balance of the building envelope does not exceed 20<br />

percent.<br />

For a partially enclosed building, the nominal value of GCpi is ±0.55. We employ the<br />

statistics from Ellingwood <strong>and</strong> Tekie (1999), assuming that the value of GCpi is normally<br />

distributed with mean = 0.46 <strong>and</strong> COV = 0.33.<br />

According to ASCE 7 st<strong>and</strong>ard (American Society of Civil Engineers, 2003), ‘opening<br />

in a wall’ is defined as any failed window, door, garage door, or roof panel. However,<br />

as indicated previously, the FPHLP model assumes that internal pressure is a function of<br />

failed windows, doors <strong>and</strong> garage doors only. Roof panels do not affect the internal pressure.<br />

In the results section, we will examine the effect of omitting the failed roof panels in the<br />

calculation of internal pressure. In addition, we will also examine the effect of an assumption<br />

made by Lee <strong>and</strong> Rosowsky (2005), in which a structure becomes partially enclosed once<br />

there is any damage on roof panels.


3.4.3 Eurocode<br />

Chapter 3. Internal Pressure in Low-Rise <strong>Structures</strong> 49<br />

In Eurocode (European Committee <strong>for</strong> St<strong>and</strong>ardization/Technical committee (CEN/TC<br />

250), 2004), the value of internal pressure depends on the areas <strong>and</strong> locations of openings<br />

on the building envelope. If the area of openings on one face is at least twice the areas<br />

of openings in the remaining faces, that face is defined as he dominant face. The updated<br />

internal pressure coefficient is a fraction of the external pressure coefficient acting on the<br />

openings on the dominant face. If a dominant face does not exist, the internal pressure<br />

coefficient is a function of (1) the ratio of the height <strong>and</strong> the depth of the building <strong>and</strong><br />

(2) the opening ratio, which is defined as the sum area of openings subjected to negative<br />

external pressure divided by the sum areas of all openings. The calculation is much more<br />

lengthy than the previous two methods <strong>and</strong> the details are not described here due to space<br />

constraint. Similar to the ASCE 7 internal pressure determination method, we will also<br />

examine the effect of omitting the failed roof panels in the calculation of updated internal<br />

pressure.<br />

For comparison, the Eurocode pressure internal coefficient (Cpi) has to be converted to<br />

the equivalent ASCE values (GCpi)eq (Oh et al., 2007):<br />

(GCpi)eq =<br />

1/2ρV 2<br />

10m,10min Ce(z)<br />

1/2ρV 2<br />

10m,3gust KztKhKdI Cpi, (3.4)<br />

where V10m,10min is the 10-min-mean wind velocity at 10 m, V10m,3gust is the 3-second gust<br />

wind speed, Ce(z) is the exposure factor in Eurocode. For consistency, we assume that<br />

(GCpi)eq has the same nominal-to-mean ratio <strong>and</strong> COV as in the ASCE internal pressure<br />

adjustment mechanism. Note that we are not using the Eurocode to calculate the wind<br />

load. We are referencing the Eurocode only regarding the internal pressure adjustment<br />

given the occurrence of building envelope damage.


Chapter 3. Internal Pressure in Low-Rise <strong>Structures</strong> 50<br />

3.5 <strong>Wind</strong>borne Debris Impact<br />

In tropical cyclones, windborne debris contributes significantly to the damage of build-<br />

ing envelope, particularly glass openings (National Institute of St<strong>and</strong>ards <strong>and</strong> Technology,<br />

2006b). For simplicity, we assume that glass windows are the only structural components<br />

vulnerable to windborne debris. Both HAZUS-MH hurricane model (Federal Emergency<br />

<strong>Management</strong> Agency, 2006) <strong>and</strong> the FPHLP Model (Gurley et al., 2005b) <strong>for</strong>mulate the<br />

probability of debris damage to an opening, PV (D), at a given wind speed (V ) as:<br />

PV (D) = 1 − exp[−λq(1 − P (ζ − ζ0))], (3.5)<br />

where λ is the mean number of missile impacts on the building, q is the fraction of the<br />

building surface covered by the opening, P (ζ − ζ0) is the probability that the momentum<br />

(ζ) of the debris is less than the damage threshold value (ζ0) given an impact.<br />

In the FPHLP Model (Gurley et al., 2005b), the parameters λ <strong>and</strong> P (ζ − ζ0) are some<br />

functions of wind speed (V ) <strong>and</strong> are expressed in five different parameters:<br />

P (a window fails due to impact) = 1 − exp(−ANABCD), (3.6)<br />

where A is the fraction of potential missile objects in the air, NA is the total number of<br />

available missle objects, B is the fraction or airborne missiles that hit the house, C is<br />

the fraction of the impact wall that is glass, <strong>and</strong> D is the probability that the impacting<br />

missiles have momentum above damage threshold. As explained in the report of Florida<br />

public hurricane loss projection model (Gurley et al., 2005b), NA is chosen as 100. A is<br />

modeled as a Gaussian cumulative density function of wind speed, with a mean value of<br />

135 mph <strong>and</strong> a st<strong>and</strong>ard deviation of 15 mph. B is a linear function of wind speed, with a<br />

value of zero at 50 mph <strong>and</strong> a value of 0.4 at 250 mph. D is a Gaussian cumulative density


Chapter 3. Internal Pressure in Low-Rise <strong>Structures</strong> 51<br />

function of wind speed with a mean of 70 mph <strong>and</strong> a st<strong>and</strong>ard deviation of 10 mph. While<br />

other advanced windborne debris models (e.g. Lin <strong>and</strong> Vanmarcke, 2010; Lin et al., 2010)<br />

are available, we employ the FPHLP debris model parameters in this study because we are<br />

considering a st<strong>and</strong>alone structure <strong>and</strong> do not consider its interaction with the surrounding<br />

environment.<br />

3.6 Damage Evaluation<br />

Given the external pressure model, the internal pressure model, the windborne debris model<br />

<strong>and</strong> the resistance model, the damage of building envelope may be evaluated at any wind<br />

speed <strong>and</strong> wind direction. The following explains the damage evaluation mechanism math-<br />

ematically.<br />

3.6.1 Component Failure Probability<br />

A binary r<strong>and</strong>om variable, Ai, may be defined to indicate the damage condition of the i-th<br />

roof panel or wall opening:<br />

⎧<br />

⎪⎨ 0 if component i has not failed<br />

Ai =<br />

⎪⎩ 1 if component i has failed<br />

(3.7)<br />

Hence, a set A = {Ai : Ai = 1, 1 ≤ i ≤ n}, where n is the number of components, can<br />

be defined to indicate the damage condition of the entire building envelope. A = ∅ means<br />

an intact builiding envelope. Given A = a, we may define the probability distribution<br />

of the internal pressure coefficient value (GCpi) using one of the internal pressure models<br />

introduced previously. Along with the probability distributions of the external pressure<br />

coefficient (GCp) <strong>and</strong> other parameters (GCpi, Kh, Kzt <strong>and</strong> Kd in equations 3.1 <strong>and</strong> 3.2,<br />

we can evaluate the conditional probability density function f Li|a(·) of the wind load (Li)<br />

acting on any structural component i. There<strong>for</strong>e, given the cumulative distribution FRi (·)


Chapter 3. Internal Pressure in Low-Rise <strong>Structures</strong> 52<br />

of the component resistance (Ri), the uplift or pressure failure probability of the component<br />

is defined as:<br />

⎧ ∞<br />

⎪⎨ FRi<br />

P (Ai = 1|A = a) = 0<br />

⎪⎩<br />

(x)fLi|a(x)dx if Ai ∈ a<br />

1 if Ai ∈ a<br />

. (3.8)<br />

The above equation only applies to roof panel, door <strong>and</strong> garage door, which fail due to<br />

wind pressure only <strong>and</strong> is not affected by windborne debris impact. For windows, we have<br />

to consider the windborne debris impact in addition to wind pressure. Assuming that both<br />

failure modes are independent of each other, the failure probability of a window is:<br />

P (Ai = 1|A = a) = P (pressure i|A = a) + P (debrisi|A = a)<br />

−P (pressure i|A = a) · P (debrisi|A = a),<br />

(3.9)<br />

where P (pressure i|A = a) is equal to the expression in equation 3.8. P (debrisi|A = a),<br />

which is the probability that the i-th window is damaged by debris impact given the damage<br />

condition A = a, may be obtained from the windborne debris model.<br />

3.6.2 System Failure Probability<br />

The system damage simulation is graphically illustrated in Figure 3.3, <strong>and</strong> is explained<br />

mathematically as follows. Given a damage condition A = ap , the probability that A<br />

becomes aq is equal to:<br />

⎧<br />

⎪⎨<br />

P (A = aq|A = ap) =<br />

⎪⎩<br />

0 if ∃k : Ak ∈ ap, Ak ∈ aq<br />

n<br />

i=0 P (Aq,i = 1|A = ap) otherwise<br />

, (3.10)<br />

where the <strong>for</strong>mulation of P (Ai = 1|A = a) is given in equations 3.8 <strong>and</strong> 3.9. The condition<br />

Ak ∈ ap <strong>and</strong> Ak ∈ aq means that once a component fails, it is always considered as failed


<strong>Wind</strong> Speed &<br />

<strong>Wind</strong> Direction<br />

Chapter 3. Internal Pressure in Low-Rise <strong>Structures</strong> 53<br />

<strong>Wind</strong>ows Only<br />

External<br />

Pressure<br />

Internal<br />

Pressure<br />

Debris<br />

Impact<br />

Roof Panels, <strong>Wind</strong>ows, Doors <strong>and</strong> Garage Doors<br />

<strong>Wind</strong><br />

Pressure<br />

if additional envelope damage occurs<br />

Figure 3.3: Flowchart of system damage simulation.<br />

<strong>and</strong> cannot be repaired during the damage simulation in the model.<br />

Envelope<br />

Damage<br />

Given an intact structure, a failure sequence over time can be defined as {∅, a1, a2, . . . , am},<br />

where ap ⊂ aq, p < q. The building envelope is intact at the beginning (A = ∅), then it<br />

becomes a1 <strong>and</strong> eventually a2 over time because of the change in internal pressure. The<br />

damage condition stops at am when it achieves the equilibrium between the internal pres-<br />

sure <strong>and</strong> the envelope damage. To calculate the occurrence probability that A = am, we<br />

have to consider all the failure sequences that lead to am:<br />

P (A = am) =<br />

<br />

P (A = a2|A = a1) · · · ·<br />

all {a1,...,am}<br />

· P (A = am|A = am−1) · P (A = am|A = am).<br />

(3.11)<br />

The last term P (A = am|A = am) in the above equation accounts <strong>for</strong> the probability<br />

that the failure sequence stops at am, at which an equilibrium has achieved between the<br />

internal pressure <strong>and</strong> building envelope damage. To avoid repeated counts of windborne<br />

debris impact after every change of internal pressure value, we only account <strong>for</strong> the debris


Chapter 3. Internal Pressure in Low-Rise <strong>Structures</strong> 54<br />

impact at the beginning when the building envelope is intact (A = ∅):<br />

P (debrisi|A = a) = 0 if a = ∅. (3.12)<br />

To quantify the damage condition of the building envelope, we can easily obtain the number<br />

of failed roof panels <strong>and</strong> the number of failed wall openings from the set A. The loss ratio is<br />

obtained by dividing the number of failed components over the total number of components.<br />

3.7 Results<br />

3.7.1 Baseline structure B under 150-mph wind<br />

Baseline structure B under 150-mph wind towards the structure’s front face is served as an<br />

example to illustrate the damage mechanism that is illustrated earlier in Figure 3.3. For<br />

simplicity, windborne debris impact is omitted. 10,000 numerical simulations are carried<br />

out <strong>for</strong> each of the three internal pressure adjustment methods that are introduced earlier.<br />

Figure 3.4 presents the interplay between the internal pressure <strong>and</strong> the building envelope<br />

damage using each of the three methods.<br />

In each simulation, two values <strong>for</strong> the initial internal pressure coefficient are r<strong>and</strong>omly<br />

selected from two normal distributions (mean = +0.15, COV = 0.33; <strong>and</strong> mean = -0.15,<br />

COV = 0.33). With the two initial internal pressure coefficients, two net pressure values are<br />

calculated <strong>for</strong> each roof panel <strong>and</strong> wall opening, <strong>and</strong> the larger one will be used to estimate<br />

the failure of the structural components. In Figure 3.4, the histogram labeled ‘internal<br />

pressure coefficient’ shows the distribution of the internal pressure coefficient which yields<br />

the larger load. The distribution of the internal pressure coefficient <strong>and</strong> the loss ratio of roof<br />

panels <strong>and</strong> wall openings are the same in every method’s step 1, because internal pressure<br />

adjustment does not apply until step 2. In most cases of step 1, the garage door, which is<br />

has the lowest resistance in the model, fail. In some cases, there may be no wall opening


(a)<br />

STEP 1<br />

STEP 2<br />

STEP 3<br />

(b)<br />

STEP 1<br />

STEP 2<br />

STEP 3<br />

(c)<br />

STEP 1<br />

STEP 2<br />

STEP 3<br />

-1.5<br />

-1.5<br />

-1.5<br />

-1.5<br />

-1.5<br />

-1.5<br />

-1.5<br />

-1.5<br />

-1.5<br />

Chapter 3. Internal Pressure in Low-Rise <strong>Structures</strong> 55<br />

Internal Pressure Coefficient Damaged Roof Panels Damaged Wall Openings<br />

0 +1.5<br />

0 +1.5<br />

0 +1.5<br />

0%<br />

0%<br />

0%<br />

100% 0%<br />

100% 0%<br />

100% 0%<br />

Internal Pressure Coefficient Damaged Roof Panels Damaged Wall Openings<br />

0 +1.5<br />

0 +1.5<br />

0 +1.5<br />

0%<br />

0%<br />

0%<br />

100% 0%<br />

100% 0%<br />

100% 0%<br />

Internal Pressure Coefficient Damaged Roof Panels Damaged Wall Openings<br />

0 +1.5<br />

0 +1.5<br />

0 +1.5<br />

0%<br />

0%<br />

100% 0%<br />

100% 0%<br />

100% 0%<br />

Figure 3.4: Distributions of internal pressure coefficient <strong>and</strong> building envelope damage<br />

of baseline structure B under a 150-mph wind towards the structure’s front face using<br />

(a) FPHLP (b) ASCE 7 St<strong>and</strong>ard (c) Eurocode internal pressure adjustment mechanism.<br />

<strong>Wind</strong>borne debris impact is not considered in the analysis. Each roof panel has the same<br />

weight in the calculation of loss ratio regardless of its area. Similarly, each window, door<br />

<strong>and</strong> garage door has the same weight.<br />

0%<br />

100%<br />

100%<br />

100%<br />

100%<br />

100%<br />

100%<br />

100%<br />

100%<br />

100%


Chapter 3. Internal Pressure in Low-Rise <strong>Structures</strong> 56<br />

failures at all, or there are several broken windows in addition of the failed garage door. At<br />

least 20% of the roof panels are uplifted in most cases of step 1. Based on the damage in<br />

step 1, the internal pressure distribution is updated in step 2 using three different internal<br />

pressure adjustment methods.<br />

Using the internal pressure adjustment method in the FPHLP model (see Figure 3.4a),<br />

the internal pressure coefficient becomes around 1.0 in many of the simulations in step 2. It<br />

is because in this model, the internal pressure coefficient is a weighted average of external<br />

pressure coefficients acting on the broken wall openings. In most cases, as a result of the<br />

failed garage door on the front face, the updated internal pressure coefficient becomes ∼1.0,<br />

which is the value of the external pressure coefficient acting on the front face. The positive<br />

internal pressure intensifies the uplift load on the roof, causing much more damage on the<br />

roof panels. It also increases the pressure loads on the walls which experience suction from<br />

leeward winds, causing more damage on the windows <strong>and</strong> doors. Based on the damage<br />

level at step 2, the internal pressure is updated again. But it virtually does not induce<br />

any further damage. The failure mechanism finishes after step 3, in which each of the<br />

10,000 simulations has achieved equilibrium between the internal pressure <strong>and</strong> the building<br />

envelope damage.<br />

In the ASCE 7 St<strong>and</strong>ard internal pressure method (see Figure 3.4b), there is no internal<br />

pressure adjustment in almost all simulations, because it is very rare that the first condition<br />

of a partially enclosed building is satisfied. When significant roof panels are uplifted, it is<br />

almost impossible that the sum area of failed windows <strong>and</strong> doors will exceed the sum area of<br />

uplifted roof panels. As a result, the total area of openings in a wall that receives positive<br />

pressure cannot exceed the sum area of openings that receive zero or negative external<br />

pressure, since the roof is always subjected to uplift (negative) pressure. This may not<br />

reflect the reality very well since it is difficult to believe that the internal pressure does not<br />

change even after most roof panels are uplifted, exposing the internal of the structure to the


Chapter 3. Internal Pressure in Low-Rise <strong>Structures</strong> 57<br />

wind. There<strong>for</strong>e, as stated in the previous section, we are modifying the ASCE 7 internal<br />

pressure mechanism so that the internal pressure depends on the wall openings only. We are<br />

also modifying the mechanism so that the building becomes partially enclosed immediately<br />

once any component fails. The results will be shown <strong>and</strong> discussed later in this chapter.<br />

In the Eurocode internal pressure method (see Figure 3.4c), the internal pressure is<br />

based on the area <strong>and</strong> location of failed wall openings <strong>and</strong> failed roof panels. The internal<br />

pressure coefficient becomes around either -1.10 or -0.25 in almost all simulations of step 2.<br />

In some simulations, the area of uplifted roof panels is at least twice the area of the failed<br />

windows <strong>and</strong> doors. According to the Eurocode, the internal pressure coefficient hence<br />

becomes a fraction of the negative external pressure coefficients that act on the roof panels,<br />

which is around -1.10. In other simulations, the internal pressure becomes around -0.25,<br />

because when the roof damage is not severe, the internal pressure is calculated based on the<br />

opening ratio, which is defined as the sum area of openings subjected to negative external<br />

pressure divided by the sum areas of all openings. Similar to the FPHLP model in Figure<br />

3.4a, after step 2, the internal pressure becomes smaller (closer to zero) in many simulations,<br />

leading to a smaller wind load acting on the structure. It is because in both methods, the<br />

internal pressure coefficient is some average of the external pressure coefficients acting on<br />

several damaged components. More damages often mean that the internal pressure is the<br />

average of more positive <strong>and</strong> negative values, <strong>and</strong> has a smaller absolute value.<br />

The results above do not consider the presence of windborne debris impact. Herein we<br />

use the same example but include windborne debris impact in the simulations. Figure 3.5<br />

presents the distributions of internal pressure coefficient <strong>and</strong> building envelope damage using<br />

each of the three internal pressure adjustment methods. Compared to the previous results,<br />

windborne debris impact leads to additional window damage in step 1 (see step 1 in Figure<br />

3.5). The additional failed windows may or may not affect the internal pressure value or the<br />

roof panel damage in step 2, depending on the method of internal pressure adjustment. In


(a)<br />

STEP 1<br />

STEP 2<br />

STEP 3<br />

(b)<br />

STEP 1<br />

STEP 2<br />

STEP 3<br />

(c)<br />

STEP 1<br />

STEP 2<br />

STEP 3<br />

-1.5<br />

-1.5<br />

-1.5<br />

-1.5<br />

-1.5<br />

-1.5<br />

-1.5<br />

-1.5<br />

-1.5<br />

Chapter 3. Internal Pressure in Low-Rise <strong>Structures</strong> 58<br />

Internal Pressure Coefficient Damaged Roof Panels Damaged Wall Openings<br />

0 +1.5<br />

0 +1.5<br />

0 +1.5<br />

0%<br />

0%<br />

0%<br />

100% 0%<br />

100% 0%<br />

100% 0%<br />

Internal Pressure Coefficient Damaged Roof Panels Damaged Wall Openings<br />

0 +1.5<br />

0 +1.5<br />

0 +1.5<br />

0%<br />

0%<br />

0%<br />

100% 0%<br />

100% 0%<br />

100% 0%<br />

Internal Pressure Coefficient Damaged Roof Panels Damaged Wall Openings<br />

0 +1.5<br />

0 +1.5<br />

0 +1.5<br />

0%<br />

0%<br />

100% 0%<br />

100% 0%<br />

100% 0%<br />

Figure 3.5: Distributions of internal pressure coefficient <strong>and</strong> building envelope damage<br />

of baseline structure B under a 150-mph wind towards the structure’s front face using<br />

(a) FPHLP (b) ASCE 7 St<strong>and</strong>ard (c) Eurocode internal pressure adjustment mechanism.<br />

<strong>Wind</strong>borne debris impact is considered in the analysis. Each roof panel has the same weight<br />

in the calculation of loss ratio regardless of its area. Similarly, each window, door <strong>and</strong> garage<br />

door has the same weight.<br />

0%<br />

100%<br />

100%<br />

100%<br />

100%<br />

100%<br />

100%<br />

100%<br />

100%<br />

100%


Chapter 3. Internal Pressure in Low-Rise <strong>Structures</strong> 59<br />

the FPHLP model, the internal pressure distribution changes significantly in the presence of<br />

windborne debris impact, because the internal pressure value is very sensitive to the number<br />

of failed wall openings (see equation 3.3). This considerably increases the roof panel loss<br />

ratio. However, in the ASCE 7 internal pressure adjustment method, the additional window<br />

damage does not make the structure a ‘partially enclosed building’. There<strong>for</strong>e, it does not<br />

affect the internal pressure calculation, leaving the roof panel loss ratio the same as in<br />

the previous results. In the Eurocode, the windborne debris impact changes the internal<br />

pressure distribution in certain extent, but such change in internal pressure does not lead<br />

to much difference in roof panel damage distribution.<br />

3.7.2 Roof Damage over 50–250 mph <strong>Wind</strong> Speed Range<br />

The results above demonstrated the behavior of baseline structure B under a sepcific wind<br />

speed <strong>and</strong> wind direction. We can exp<strong>and</strong> the analysis by running the same type of sim-<br />

ulations using different baseline structures, wind speeds <strong>and</strong> directions, <strong>and</strong> more internal<br />

pressure adjustment mechanisms. The resultant roof vulnerability curves are presented in<br />

Figure 3.6. Herein, the roof vulnerability is defined as the mean loss ratio of roof panels<br />

at the last step of the internal pressure adjustment process, which is step 3 in Figures 3.4<br />

<strong>and</strong> 3.5. Since the results are the same <strong>for</strong> both wind directions parallel to ridgeline, we are<br />

presenting only one of them. There<strong>for</strong>e, the figure shows the results of three wind angles,<br />

instead of the four cardinal wind directions. Also, different assumptions within the ASCE<br />

7 <strong>and</strong> Eurocode internal pressure adjustment methods are considered, leading to a total<br />

of seven different internal pressure adjustment models. In the figure, ‘(instant)’ in ASCE<br />

7 means that a building becomes ‘partially enclosed’ immediately after occurrence of any<br />

degree of damage on the building envelope.<br />

As wind speed increases, mean roof damage increases nonlinearly in almost all cases.<br />

For each wind direction (presented in each subfigure (a), (b) <strong>and</strong> (c) in Figure 3.6), if


(a)<br />

WITHOUT WINDBORNE DEBRIS WITH WINDBORNE DEBRIS<br />

(b)<br />

Mean Loss Ratio<br />

WITHOUT WINDBORNE DEBRIS WITH WINDBORNE DEBRIS<br />

Mean Loss Ratio<br />

Mean Loss Ratio<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

Chapter 3. Internal Pressure in Low-Rise <strong>Structures</strong> 60<br />

BASELINE STRUCTURE A BASELINE STRUCTURE B BASELINE STRUCTURE C<br />

0<br />

50 100 150 200 250<br />

<strong>Wind</strong> Speed (mph)<br />

0<br />

50 100 150 200 250<br />

<strong>Wind</strong> Speed (mph)<br />

0.6<br />

0.6<br />

BASELINE STRUCTURE B BASELINE 0.8 STRUCTURE C<br />

0.8<br />

0.8<br />

Mean Loss Ratio<br />

Mean Loss Ratio<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

Mean Loss Ratio<br />

Mean Loss Ratio<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

50 100 150 200 250<br />

<strong>Wind</strong> Speed (mph)<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

50 100 150 200 250<br />

<strong>Wind</strong> Speed (mph)<br />

Mean Loss Ratio<br />

Mean Loss Ratio<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

50 100 150 200 250<br />

<strong>Wind</strong> Speed (mph)<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

50 100 150 200 250<br />

<strong>Wind</strong> Speed (mph)<br />

BASELINE STRUCTURE A BASELINE STRUCTURE B BASELINE STRUCTURE C<br />

0.2<br />

0.2<br />

0.2<br />

BASELINE STRUCTURE A BASELINE STRUCTURE B BASELINE STRUCTURE C<br />

1<br />

0<br />

50<br />

1<br />

100 150 200 250<br />

0<br />

50<br />

1<br />

100 150 200 250<br />

LINE STRUCTURE A<br />

0.8<br />

1<br />

<strong>Wind</strong> Speed (mph)<br />

<strong>Wind</strong> Speed (mph)<br />

BASELINE STRUCTURE B BASELINE STRUCTURE C<br />

0.8<br />

0.8<br />

1<br />

1<br />

1<br />

WITHOUT WINDBORNE DEBRIS WITH WINDBORNE DEBRIS<br />

Mean Loss Ratio<br />

0<br />

50 100 0.2 150 200 250<br />

<strong>Wind</strong> Speed (mph)<br />

0<br />

00 250 50 100 150 200 250<br />

<strong>Wind</strong> Speed (mph)<br />

Mean Loss Ratio<br />

Mean Loss Ratio<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.8<br />

Mean Loss Ratio<br />

Mean Loss Ratio<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0<br />

50 100 150 200 250<br />

<strong>Wind</strong> Speed (mph)<br />

1<br />

URE B<br />

0.4<br />

1<br />

BASELINE 0.6 STRUCTURE C<br />

0.6<br />

0.4<br />

0.6<br />

0.6<br />

0.4<br />

0.6<br />

LEGEND<br />

.8<br />

0.2<br />

0.8<br />

0.2<br />

0.2<br />

1<br />

0.4<br />

0.4<br />

0.4<br />

0.4<br />

0.4<br />

LEGEND<br />

.6<br />

0<br />

0.6<br />

0<br />

0<br />

0.8 50 100 150 200 0.2 250 50 100 150 200 0.2 250 50 100 150 200 0.2 250<br />

0.2<br />

<strong>Wind</strong> Speed (mph)<br />

0.2<br />

No internal pressure<br />

<strong>Wind</strong> Speed (mph)<br />

<strong>Wind</strong> Speed (mph)<br />

adjustment<br />

.4<br />

0.4<br />

0.6<br />

0<br />

LEGEND 0<br />

FLPHP<br />

1<br />

0<br />

0<br />

50<br />

1<br />

100 1500 200 250 50<br />

1<br />

100 150 200 250 50<br />

150 200 250 50 100 150 200 250 50 100 150 200 250<br />

ASCE 7<br />

.2<br />

<strong>Wind</strong> Speed (mph)<br />

<strong>Wind</strong> Speed (mph)<br />

ind Speed (mph)<br />

0.2<br />

0.4<br />

<strong>Wind</strong> Speed (mph)<br />

<strong>Wind</strong> Speed<br />

No internal<br />

(mph)<br />

pressure<br />

(excl. failed roof panels)<br />

0.8<br />

0.8 LEGEND<br />

adjustment 0.8<br />

ASCE 7<br />

FLPHP<br />

(incl. failed roof panels)<br />

0<br />

1<br />

0<br />

1<br />

50 100 0.2 150<br />

No internal pressure<br />

0.6 200 250 50LEGEND 100 150 0.6 200 250<br />

ASCE 7<br />

ASCE 7 (instant)<br />

adjustment<br />

0.6<br />

<strong>Wind</strong> Speed (mph)<br />

<strong>Wind</strong> Speed (mph)<br />

(excl. failed roof panels)<br />

(excl. failed roof panels)<br />

0.8<br />

FLPHP 0.8<br />

ASCE 7<br />

ASCE 7 (instant)<br />

0<br />

00 250 50 0.4 100 150 200 250 0.4 ASCE 7<br />

(incl. failed roof<br />

0.4<br />

panels)<br />

(incl. failed roof panels)<br />

1<br />

1<br />

<strong>Wind</strong> Speed (mph)<br />

(excl. failed roof panels) ASCE 7 (instant)<br />

EUROCODE<br />

0.6<br />

ASCE 7<br />

0.6<br />

(excl. failed roof panels)<br />

(excl. failed roof panels)<br />

.8<br />

0.2<br />

0.8<br />

0.2 (incl. failed roof panels) ASCE 7 (instant) 0.2<br />

EUROCODE<br />

1<br />

ASCE 7 (instant)<br />

(incl. failed roof panels)<br />

(incl. failed roof panels)<br />

0.4<br />

0.4<br />

(excl. failed roof panels) EUROCODE<br />

.6<br />

0<br />

0.6<br />

0<br />

0.8 50 100 150 200 250 50ASCE 7 100 (instant)<br />

(excl. failed roof 0panels)<br />

150 200 250 50 100 150 200 250<br />

0.2<br />

<strong>Wind</strong> Speed (mph)<br />

(incl. failed 0.2<br />

<strong>Wind</strong><br />

roof<br />

Speed<br />

panels)<br />

(mph) Figure EUROCODE 3.6: To be<strong>Wind</strong> continued.<br />

Speed (mph)<br />

.4<br />

0.4<br />

EUROCODE<br />

(incl. failed roof panels)<br />

0.6<br />

(excl. failed roof panels)<br />

0<br />

150 200 250 50 100 150 200 EUROCODE<br />

0<br />

250 50 100 150 200 250<br />

.2ind<br />

Speed (mph)<br />

0.2<br />

(incl. failed roof panels)<br />

0.4<br />

<strong>Wind</strong> Speed (mph)<br />

<strong>Wind</strong> Speed (mph)<br />

No internal pressure<br />

adjustment<br />

FLPHP<br />

ASCE 7<br />

(excl. failed roof panels)<br />

ASCE 7<br />

(incl. failed roof panels)<br />

100 ASCE 1507 (instant) 200 250<br />

<strong>Wind</strong> (excl. Speed failed (mph) roof panels)<br />

ASCE 7 (instant)<br />

(incl. failed roof panels)<br />

EUROCODE<br />

(excl. failed roof panels)<br />

EUROCODE<br />

(incl. failed roof panels)<br />

Mean Loss Ratio<br />

Mean Loss Ratio<br />

Mean Loss Ratio<br />

Mean Loss Ratio<br />

Mean Loss Ratio<br />

Mean Loss Ratio<br />

Mean Loss Ratio<br />

Mean Loss Ratio<br />

Mean Loss Ratio<br />

Mean Loss Ratio<br />

0<br />

50 100 150 200 250<br />

<strong>Wind</strong> Speed (mph)<br />

Mean Loss Ratio<br />

Mean Loss Ratio<br />

0.6<br />

1<br />

0.8<br />

LEGEND<br />

No internal pressure<br />

adjustment<br />

FLPHP<br />

ASCE 7<br />

(excl. failed roof panels)<br />

ASCE 7<br />

(incl. failed roof panels)<br />

ASCE 7 (instant)<br />

(excl. failed roof panels)<br />

ASCE 7 (instant)<br />

(incl. failed roof panels)<br />

EUROCODE<br />

(excl. failed roof panels)<br />

EUROCODE<br />

(incl. failed roof panels)<br />

LEGEND<br />

No internal pressure<br />

adjustment<br />

FLPHP<br />

ASCE 7<br />

(excl. failed roof panels)<br />

ASCE 7<br />

(incl. failed roof panels)<br />

ASCE 7 (instant)<br />

(excl. failed roof panels)<br />

ASCE 7 (instant)<br />

(incl. failed roof panels)<br />

EUROCODE<br />

(excl. failed roof panels)<br />

EUROCODE<br />

(incl. failed roof panels)


Mean Loss Ratio<br />

WITHOUT WINDBORNE DEBRIS WITH WINDBORNE DEBRIS<br />

0.6<br />

0.6<br />

BASELINE STRUCTURE B BASELINE 0.8 STRUCTURE C<br />

0.8<br />

0.8<br />

Mean Loss Ratio<br />

Mean Loss Ratio<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

Chapter 3. Internal Pressure in Low-Rise <strong>Structures</strong> 61<br />

BASELINE STRUCTURE A BASELINE STRUCTURE B BASELINE STRUCTURE C<br />

0.2<br />

0.2<br />

0.2<br />

BASELINE STRUCTURE A BASELINE STRUCTURE B BASELINE STRUCTURE C<br />

1<br />

0<br />

50<br />

1<br />

100 150 200 250<br />

0<br />

50<br />

1<br />

100 150 200 250<br />

LINE STRUCTURE A<br />

0.8<br />

1<br />

<strong>Wind</strong> Speed (mph)<br />

<strong>Wind</strong> Speed (mph)<br />

BASELINE STRUCTURE B BASELINE STRUCTURE C<br />

0.8<br />

0.8<br />

1<br />

1<br />

1<br />

WITHOUT WINDBORNE DEBRIS WITH WINDBORNE DEBRIS<br />

Mean Loss Ratio<br />

0<br />

50 100 0.2 150 200 250<br />

<strong>Wind</strong> Speed (mph)<br />

0<br />

00 250 50 100 150 200 250<br />

<strong>Wind</strong> Speed (mph)<br />

Mean Loss Ratio<br />

Mean Loss Ratio<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.8<br />

Mean Loss Ratio<br />

Mean Loss Ratio<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0<br />

50 100 150 200 250<br />

<strong>Wind</strong> Speed (mph)<br />

1<br />

URE B<br />

0.4<br />

1<br />

BASELINE 0.6 STRUCTURE C<br />

0.6<br />

0.4<br />

0.6<br />

0.6<br />

0.4<br />

0.6<br />

LEGEND<br />

.8<br />

0.2<br />

0.8<br />

0.2<br />

0.2<br />

1<br />

0.4<br />

0.4<br />

0.4<br />

0.4<br />

0.4<br />

LEGEND<br />

.6<br />

0<br />

0.6<br />

0<br />

0<br />

0.8 50 100 150 200 0.2 250 50 100 150 200 0.2 250 50 100 150 200 0.2 250<br />

0.2<br />

<strong>Wind</strong> Speed (mph)<br />

0.2<br />

No internal pressure<br />

<strong>Wind</strong> Speed (mph)<br />

<strong>Wind</strong> Speed (mph)<br />

adjustment<br />

.4<br />

0.4<br />

0.6<br />

0<br />

LEGEND 0<br />

FLPHP<br />

1<br />

0<br />

0<br />

50<br />

1<br />

100 1500 200 250 50<br />

1<br />

100 150 200 250 50<br />

150 200 250 50 100 150 200 250 50 100 150 200 250<br />

ASCE 7<br />

.2<br />

<strong>Wind</strong> Speed (mph)<br />

<strong>Wind</strong> Speed (mph)<br />

ind Speed (mph)<br />

0.2<br />

0.4<br />

<strong>Wind</strong> Speed (mph)<br />

<strong>Wind</strong> Speed<br />

No internal<br />

(mph)<br />

pressure<br />

(excl. failed roof panels)<br />

0.8<br />

0.8 LEGEND<br />

adjustment 0.8<br />

ASCE 7<br />

FLPHP<br />

(incl. failed roof panels)<br />

0<br />

1<br />

0<br />

1<br />

50 100 0.2 150<br />

No internal pressure<br />

0.6 200 250 50LEGEND 100 150 0.6 200 250<br />

ASCE 7<br />

ASCE 7 (instant)<br />

adjustment<br />

0.6<br />

<strong>Wind</strong> Speed (mph)<br />

<strong>Wind</strong> Speed (mph)<br />

(excl. failed roof panels)<br />

(excl. failed roof panels)<br />

0.8<br />

FLPHP 0.8<br />

ASCE 7<br />

ASCE 7 (instant)<br />

0<br />

00 250 50 0.4 100 150 200 250 0.4 ASCE 7<br />

(incl. failed roof<br />

0.4<br />

panels)<br />

(incl. failed roof panels)<br />

1<br />

1<br />

<strong>Wind</strong> Speed (mph)<br />

(excl. failed roof panels) ASCE 7 (instant)<br />

EUROCODE<br />

0.6<br />

ASCE 7<br />

0.6<br />

(excl. failed roof panels)<br />

(excl. failed roof panels)<br />

.8<br />

0.2<br />

0.8<br />

0.2 (incl. failed roof panels) ASCE 7 (instant) 0.2<br />

EUROCODE<br />

1<br />

ASCE 7 (instant)<br />

(incl. failed roof panels)<br />

(incl. failed roof panels)<br />

0.4<br />

0.4<br />

(excl. failed roof panels) EUROCODE<br />

.6<br />

0<br />

0.6<br />

0<br />

0.8 50 100 150 200 250 50ASCE 7 100 (instant)<br />

(excl. failed roof 0panels)<br />

150 200 250 50 100 150 200 250<br />

0.2<br />

<strong>Wind</strong> Speed (mph)<br />

(incl. failed 0.2<br />

<strong>Wind</strong><br />

roof<br />

Speed<br />

panels)<br />

(mph)<br />

EUROCODE<br />

<strong>Wind</strong> Speed (mph)<br />

.4<br />

0.4<br />

EUROCODE<br />

(incl. failed roof panels)<br />

0.6<br />

(excl. failed roof panels)<br />

0<br />

150 200 250 50 100 150 200 EUROCODE<br />

0<br />

250 50 100 150 200 250<br />

.2ind<br />

Speed (mph)<br />

0.2<br />

(incl. failed roof panels)<br />

0.4<br />

<strong>Wind</strong> Speed (mph)<br />

<strong>Wind</strong> Speed (mph)<br />

No internal pressure<br />

adjustment<br />

FLPHP<br />

ASCE 7<br />

(excl. failed roof panels)<br />

ASCE 7<br />

(incl. failed roof panels)<br />

100 ASCE 1507 (instant) 200 250<br />

<strong>Wind</strong> (excl. Speed failed (mph) roof panels)<br />

ASCE 7 (instant)<br />

(incl. failed roof panels)<br />

EUROCODE<br />

(excl. failed roof panels)<br />

EUROCODE<br />

(incl. failed roof panels)<br />

Mean Loss Ratio<br />

Mean Loss Ratio<br />

Mean Loss Ratio<br />

Mean Loss Ratio<br />

Mean Loss Ratio<br />

Mean Loss Ratio<br />

(c)<br />

Mean Loss Ratio<br />

Mean Loss Ratio<br />

Mean Loss Ratio<br />

Mean Loss Ratio<br />

0<br />

50 100 150 200 250<br />

<strong>Wind</strong> Speed (mph)<br />

Mean Loss Ratio<br />

Mean Loss Ratio<br />

0.6<br />

Figure 3.6: Roof panel vulnerability curves: (a) <strong>Wind</strong> towards the front face of structure<br />

(b) <strong>Wind</strong> parallel to ridgeline (c) <strong>Wind</strong> towards to the back face of structure. In the legend,<br />

‘excl. failed roof panels’ means that the roof damage condition is not considered in the<br />

calculation of internal pressure. Similarly, ‘incl. failed roof panels’ means that the roof<br />

damage condition is included in the calculation of internal pressure.<br />

1<br />

0.8<br />

LEGEND<br />

No internal pressure<br />

adjustment<br />

FLPHP<br />

ASCE 7<br />

(excl. failed roof panels)<br />

ASCE 7<br />

(incl. failed roof panels)<br />

ASCE 7 (instant)<br />

(excl. failed roof panels)<br />

ASCE 7 (instant)<br />

(incl. failed roof panels)<br />

EUROCODE<br />

(excl. failed roof panels)<br />

EUROCODE<br />

(incl. failed roof panels)


Chapter 3. Internal Pressure in Low-Rise <strong>Structures</strong> 62<br />

there is no internal pressure adjustment, the roof vulnerability curves are exactly the same<br />

regardless of the choice of baseline structure or the presence of windborne debris, because<br />

all the baseline structures have the same roof <strong>and</strong> the windborne debris does not damage<br />

the roof given the setup of the model. On average, wind parallel to ridgeline is causing a<br />

smaller roof damage, because about half of the roof is well shielded <strong>and</strong> is not subjected to<br />

any load, as shown in Figure 3.2. In all cases, internal pressure adjustment increases or has<br />

no effect on the mean roof damage. The degree of such increase depends on the wind speed,<br />

the wind direction, the existence of windborne debris <strong>and</strong> the number of windows on the<br />

structure. There is no simple linear relationship between the degree of increase <strong>and</strong> any of<br />

the above parameters. The following presents a breakdown of the effect of each parameter<br />

on the mean roof panel loss ratio in each of the internal pressure adjustment methods.<br />

3.7.2.1 Percentage increase in mean roof damage ratio<br />

First, Figure 3.7 shows the range of percentage increase as well as the average percentage<br />

increase in mean roof damage, using different internal pressure adjustment methods com-<br />

paring no internal pressure adjustment over all 18 cases in Figure 3.6. The FLPHP internal<br />

pressure adjustment model has the largest overall impact on the mean roof damage, partic-<br />

ularly over the 120-130 mph wind speed (>200%). On average it contributes an addittional<br />

∼50% to the mean roof damage ratio.<br />

Comparatively, the ASCE 7 method has a relatively small effect on the mean roof<br />

damage. If we do not consider any failed roof panels in the internal pressure adjustment,<br />

the maximum percentage increase is about 70%. But the average is less than 20% over all<br />

wind speeds. There is almost no effect on the mean roof damage if we do consider the failed<br />

roof panels in the internal pressure adjustment. As explained in last section, the reason is<br />

because the condition of ‘partially enclosed’ building is difficult to satisfy. However, if it<br />

is assumed that a building becomes ‘partially enclosed’ immediately after any damage (see


Maximum Percentage<br />

Increase in<br />

Mean Roof Damage (%)<br />

Maximum Percentage Difference (%)<br />

Maximum Percentage<br />

Increase in<br />

Mean Roof Damage (%)<br />

Maximum Percentage Difference (%)<br />

Maximum Percentage<br />

Increase in<br />

Mean Roof Damage (%)<br />

Maximum Percentage Difference (%)<br />

250<br />

200<br />

150<br />

100<br />

50<br />

Chapter 3. Internal Pressure in Low-Rise <strong>Structures</strong> 63<br />

0<br />

100 150 200 250<br />

<strong>Wind</strong> Speed (mph)<br />

<strong>Wind</strong> Speed (mph)<br />

250<br />

200<br />

150<br />

100<br />

50<br />

0<br />

100 150 200 250<br />

<strong>Wind</strong> Speed (mph)<br />

<strong>Wind</strong> Speed (mph)<br />

250<br />

200<br />

150<br />

100<br />

50<br />

!"#$"<br />

%&'()(*%+,-./01+2340-5+6772+839-0:<br />

%&'()(*%+,49/01+2340-5+6772+839-0:<br />

0<br />

100 150 200 250<br />

<strong>Wind</strong> Speed (mph) (mph)<br />

Maximum Percentage<br />

Increase in<br />

Mean Roof Damage (%)<br />

Maximum Percentage Difference (%)<br />

Maximum Percentage<br />

Increase in<br />

Mean Roof Damage (%)<br />

Maximum Percentage Difference (%)<br />

Maximum Percentage<br />

Increase in<br />

Mean Maximum Roof Percentage Damage Difference (%) (%)<br />

Maximum Percentage<br />

Increase in<br />

Mean Maximum Roof Percentage Damage Difference (%) (%)<br />

250<br />

200<br />

150<br />

100<br />

50<br />

0<br />

100 150 200 250<br />

<strong>Wind</strong> Speed (mph)<br />

<strong>Wind</strong> Speed (mph)<br />

250<br />

200<br />

150<br />

100<br />

50<br />

0<br />

100 150 200<br />

<strong>Wind</strong> Speed (mph)<br />

<strong>Wind</strong> Speed (mph)<br />

250<br />

250<br />

200<br />

150<br />

100<br />

50<br />

0<br />

100 150 200 250<br />

<strong>Wind</strong> Speed (mph) (mph)<br />

250<br />

200<br />

150<br />

100<br />

50<br />

;


Chapter 3. Internal Pressure in Low-Rise <strong>Structures</strong> 64<br />

the figures labeled ASCE 7 (instant)), the effect of internal pressure on the roof damage is<br />

significantly increased. This significant increase is the same no matter whether the failed<br />

roof panels are taken into account, since in almost every simulation with wind speed over<br />

100 mph, both roof panels <strong>and</strong> wall openings almost cannot avoid damage <strong>and</strong> the building<br />

turns to ‘partially enclosed’. The internal pressure coefficient hence becomes the prescribed<br />

value <strong>for</strong> ‘partially enclosed’ building. The percentage increase is more significant when the<br />

wind speed is around 100 mph. The increases diminishes as the wind speed increases.<br />

Lastly, in the Eurocode internal pressure adjustment mechanism, if failed roof panels<br />

are omitted, the mean roof damage significantly increases in cases in which the wind speed<br />

is under 160 mph. The effect diminishes when the wind speed increases. If failed roof panels<br />

are considered in the calculation of internal pressure, the results are opposite. The increase<br />

in mean roof panel loss ratio is relatively small when the wind speed is under 160 mph, <strong>and</strong><br />

the increase can reach up to 80% as wind speed increases.<br />

3.7.2.2 Number of windows <strong>and</strong> windborne debris<br />

Second, we examine the effects of the number of windows <strong>and</strong> the windborne debris on the<br />

roof damage in the presence of internal pressure adjustment. Average roof vulnerability<br />

curves are first obtained by averaging the respective roof vulnerability curves from the<br />

four cardinal wind directions. The average roof vulnerability curve of baseline structure C<br />

without the presence of windborne debris is selected as the control that is to be compared<br />

with all other average roof vulnerability curves. The percentage differences between all the<br />

average vulnerability curves <strong>and</strong> the structure-C-without-debris average vulnerability curve<br />

are presented in Figure 3.8.<br />

The left of Figure 3.8 illustrates how the number of windows affects the mean roof<br />

damage if windborne debris impact is not accounted <strong>for</strong>. In the figure, the dotted line is<br />

always at 0%, since it represents the percentage difference between the mean roof damage


Chapter 3. Internal Pressure in Low-Rise <strong>Structures</strong> 65<br />

baseline structure C <strong>and</strong> itself. The other two lines which represent baseline structure<br />

A (which has the most windows) <strong>and</strong> baseline structure B (which has the second most<br />

windows) vary with the wind speed <strong>and</strong> the choice of internal pressure method. As indicated<br />

in the Figure, more windows do not necessarily imply a ‘worse’ internal pressure, higher<br />

wind load of the roof <strong>and</strong> hence a higher damage on roof. For example, when wind speed is<br />

around 200 mph, baseline structures A <strong>and</strong> B have a mean roof damage at least 5% lower<br />

than baseline structure C, which has much fewer windows. Also, at some wind speeds,<br />

baseline structure B, which has the second least windows, has the highest roof damage.<br />

It shows that there can be a nonlinear relationship between the number of windows <strong>and</strong><br />

the roof damage. In most of the cases, additional windows do not lead to more than 10%<br />

change in the average roof damage over the four wind angles if windborne debris impact is<br />

omitted.<br />

The right of Figure 3.8 illustrates how the number of windows affects the mean roof<br />

damage if windborne debris is considered in the model. The windborne debris impact has<br />

very significant effects on the roof damage since the percentage differences are significantly<br />

larger in all cases than the left panels, except in the cases of ASCE 7 where the structure<br />

becomes partially enclosed immediately given any building envelope damage (as indicated<br />

by ASCE (instant) in the Figure). It is because once the structure becomes partially en-<br />

closed, the internal pressure coefficient becomes the prescribed values <strong>for</strong> partially enclosed<br />

structure. Any additional damage on the building envelope can no longer affect the internal<br />

pressure. Besides, the effect of the number of windows on the roof damage is more signifi-<br />

cant in the presence of windborne debris impact, as shown by a larger percentage difference<br />

between the lines. But similar to the left panel of the figure, additional windows solely does<br />

not imply an increase or decrease in roof damage.


FPHLP<br />

ASCE 7<br />

(excl. failed<br />

roof panels)<br />

ASCE 7<br />

(incl. failed<br />

roof panels)<br />

ASCE 7<br />

(instant)<br />

(excl. failed<br />

roof panels)<br />

ASCE 7<br />

(instant)<br />

(incl. failed<br />

roof panels)<br />

Eurcode<br />

(excl. failed<br />

roof panels)<br />

Eurocode<br />

(incl. failed<br />

roof panels)<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

10 %<br />

0 %<br />

Chapter 3. Internal Pressure in Low-Rise <strong>Structures</strong> 66<br />

10%<br />

100<br />

4 %<br />

150 200 250<br />

0 %<br />

4%<br />

100<br />

3 %<br />

150 200 250<br />

0 %<br />

3%<br />

100<br />

4 %<br />

150 200 250<br />

0 %<br />

4%<br />

100<br />

4 %<br />

150 200 250<br />

0 %<br />

4%<br />

100<br />

10 %<br />

150 200 250<br />

0 %<br />

10%<br />

100<br />

30 %<br />

150 200 250<br />

0 %<br />

30%<br />

100 150 200 250<br />

LEGEND<br />

WITHOUT WINDBORNE DEBRIS WITH WINDBORNE DEBRIS<br />

150 %<br />

0 %<br />

150%<br />

100<br />

60 %<br />

150 200 250<br />

0 %<br />

60%<br />

100<br />

10 %<br />

150 200 250<br />

0 %<br />

10%<br />

100<br />

4 %<br />

150 200 250<br />

0 %<br />

4%<br />

100<br />

4 %<br />

150 200 250<br />

0 %<br />

4%<br />

100<br />

100 %<br />

150 200 250<br />

0 %<br />

100%<br />

100<br />

40 %<br />

150 200 250<br />

0 %<br />

40%<br />

100 150 200 250<br />

<strong>Wind</strong> speed (mph) <strong>Wind</strong> speed (mph)<br />

1<br />

Baseline structure A (15 windows)<br />

(without windborne debris<br />

0.8<br />

impact)<br />

Baseline structure B (11 windows)<br />

(without windborne debris impact)<br />

Baseline structure C (4<br />

0.6<br />

windows)<br />

(without windborne debris impact)<br />

0.4<br />

0.2<br />

0.2<br />

0.4<br />

0.6<br />

0.8<br />

LEGEND<br />

Baseline structure A (15 windows)<br />

(with windborne debris impact)<br />

Baseline structure B (11 windows)<br />

(with windborne debris impact)<br />

Baseline structure C (4 windows)<br />

(with windborne debris impact)<br />

Figure 3.8: Percentage differences between the average roof damages over four cardinal wind<br />

0 directions <strong>and</strong> the average roof damage of0 baseline structure C without windborne debris.<br />

0.2<br />

0.4<br />

0.6<br />

0.8<br />

1<br />

1<br />

1 0.8 0.6 0.4 0.2 0 0.2 0.4 0.6 1 0.8 0.8 1 0.6 0.4 0.2 0 0.2 0.4 0.6 0.8 1


3.8 Discussion<br />

Chapter 3. Internal Pressure in Low-Rise <strong>Structures</strong> 67<br />

As shown in the results, the internal pressure plays a very significant role in the building<br />

envelope damage over a wide range of wind speed <strong>and</strong> wind direction. In some cases, the<br />

number of failed roof panels is increased by a few times if internal pressure adjustment<br />

is implemented. However, the effect of internal pressure on the building damage very<br />

much depends on the setup of the internal pressure adjustment model. Un<strong>for</strong>tunately,<br />

there are not many design guidelines or experimental data on the simulation of internal<br />

pressure changes during the structural damage process. In fact, this may be a difficult<br />

task to introduce a simple model. As shown in past research, internal pressure depends<br />

on a series of factors including geometries of openings, overall building leakage, internal<br />

volume, flexibility of building envelope, etc. Damages on building envelope would further<br />

complicate the problem since it involves complex aerodynamics depending on the details of<br />

structural failure. But given the large effect of internal pressure on the building envelope as<br />

shown above, this is important <strong>for</strong> the wind engineering field to tackle the issue. Also, as<br />

more research ef<strong>for</strong>ts address the problem of windborne debris impact, there is an increasing<br />

dem<strong>and</strong> <strong>for</strong> the internal pressure research, since windborne debris impact on the building<br />

envelope is very closely related to the internal pressure problem.<br />

In the results, we only demonstrated the structural damage under individual wind direc-<br />

tions (wind parallel with or perpendicular to ridgeline). However, structural design requires<br />

an examination of the ‘worst’ structural behavior under a range of wind directions. Con-<br />

ventionally, some codes <strong>and</strong> st<strong>and</strong>ards (e.g. ASCE 7 St<strong>and</strong>ard) do not take into account<br />

the effect of different wind directions individually. Instead, they conservatively consider the<br />

‘worst’ wind angle <strong>for</strong> each envelope surface, evaluating the maximum wind pressure (in<br />

both positive <strong>and</strong> negative signs) over all wind angles acting on each individual structural<br />

component. The maximum load is then converted to the design wind pressure by some<br />

factor (e.g. wind directionality factor (Kd) in ASCE 7 St<strong>and</strong>ard), which accounts <strong>for</strong> the


Chapter 3. Internal Pressure in Low-Rise <strong>Structures</strong> 68<br />

probability that the maximum wind may not impact the component in its weakest orien-<br />

tation. However, if progressive damage (e.g. progressive building envelope damage due to<br />

internal pressure adjustment) is considered, it is necessary to consider the structural behav-<br />

ior under each individual wind angle in order to accurately model the initial damage <strong>and</strong><br />

hence the failure sequence. Looking at the worst-case scenario of each individual component<br />

separately <strong>and</strong> not inspecting all structural components simultaneously under a single wind<br />

direction, the simplified approach in ASCE 7 St<strong>and</strong>ard is unable to simulate the realistic<br />

failure sequence.<br />

3.9 Closing Remarks<br />

A structural model is developed in this study to examine the effect of internal pressure<br />

on the damage of building envelope. A number of internal pressure adjustment methods<br />

(FPHLP model, ASCE 7 St<strong>and</strong>ard <strong>and</strong> Eurocode) are implemented in the model to simulate<br />

the interplay between the building component failure <strong>and</strong> the internal pressure, creating a<br />

progressive damage mechanism. The mechanism is explained in details in mathematics <strong>and</strong><br />

is illustrated through a few numerical examples. The roof damage results using different<br />

baseline structures, internal pressure adjustment methods <strong>and</strong> windborne debris parameters<br />

over a wide range of wind speeds are presented. The effect of internal pressure on the roof<br />

damage is very significant, but the degree of the effect depends on the setup of the internal<br />

pressure adjustment model. It is necessary to further examine the internal pressure change<br />

in various building envelope damage conditions.


Chapter<br />

4<br />

Integrated Multi-Structural<br />

Vulnerability Model<br />

4.1 Introduction<br />

<strong>Wind</strong> pressure, a major cause of hurricane-induced damage to structures, has been studied<br />

extensively through wind tunnel experiments <strong>and</strong> in situ measurements, <strong>and</strong> approxima-<br />

tions of wind pressures acting on (components of) structures are incorporated into current<br />

vulnerability models, which include the FEMA HAZUS-MH model (Vickery et al., 2006b)<br />

<strong>and</strong> the Florida Public Hurricane Loss Projection (FPHLP) model (Gurley et al., 2006).<br />

Apart from wind pressure, windborne debris is another major cause of wind damage. Often<br />

generated from building materials damaged by high local wind pressure, windborne debris<br />

can penetrate building envelopes <strong>and</strong> induce internal pressurization, which increases wind<br />

pressure damage, generating more debris <strong>and</strong> further debris damage, amounting to a chain<br />

reaction of failures. Most current vulnerability models, including the HAZUS-MH model<br />

<strong>and</strong> the FPHLP model, consider only the (one-way) effect of debris damage on pressure<br />

damage <strong>and</strong> approximate the debris impact risk to a structure by considering only sim-<br />

ple, idealized building environments. An integrated vulnerability model is developed to<br />

69


Chapter 4. Integrated Multi-Structural Vulnerability Model 70<br />

explicitly account, <strong>for</strong> the first time, <strong>for</strong> the interaction between pressure damage <strong>and</strong> de-<br />

bris damage by fully coupling a debris risk model (Lin <strong>and</strong> Vanmarcke, 2008, 2010) <strong>and</strong><br />

a component-based pressure damage model (developed starting from the FPHLP model).<br />

This integrated methodology can be applied to general site- <strong>and</strong> storm-specific analyses <strong>for</strong><br />

residential developments consisting of (realistically) large numbers of buildings.<br />

Using the model by Lin et al. (2010), this chapter (1) illustrates how a structure is af-<br />

fected by neighboring buildings resistance against high wind; (2) contrasts the difference in<br />

damage predictions between time series analysis <strong>and</strong> point-in-time analysis during a storm<br />

passage; <strong>and</strong> (3) extends the estimation of wind-induced losses from structural components<br />

to the entire structural system, the interior system, <strong>and</strong> the utility system (including elec-<br />

trical, pluming, heating systems) of buildings. In what follows, the theory underlying the<br />

integrated pressure-debris risk model is briefly outlined. Additional details about the wind<br />

characteristics, the internal pressure model <strong>and</strong> the approximate economic loss model that<br />

are not covered by Lin et al. (2010) are described. In the end, four numerical examples of<br />

wind-damage estimation in residential developments are presented <strong>and</strong> discussed.<br />

4.2 Methodology<br />

4.2.1 <strong>Wind</strong> Characteristics<br />

An appropriate description of wind conditions is necessary to evaluate the structural damage<br />

<strong>and</strong> resultant economic loss <strong>for</strong> the defined cluster of buildings. Traditionally, though wind<br />

direction plays an important role in the hurricane damage of a structure, wind damage<br />

is often expressed as a function of wind speed only. One of the reasons is that there is<br />

very limited data about the wind direction with respect to the orientation of a residential<br />

structure in the insurance database. To accurately assess the wind risk of a structure, the<br />

damage of a structure is analyzed at eight nominal wind directions at 45 degree interval as


Chapter 4. Integrated Multi-Structural Vulnerability Model 71<br />

shown in Figure 4.1. Damage at any other wind directions may be linear interpolated from<br />

the damages at the two closest nominal wind directions.<br />

Figure 4.1: Eight nominal angles of wind incidence.<br />

During a tropical cyclone event, a structure is exposed to winds from different directions<br />

as the storm moves along its track. The greater the variation in wind directionality, the<br />

wider the structures angle of exposure to windward wind. For instance, more damage<br />

can be caused by wind-borne debris if wind directionality changes more, as debris mainly<br />

impacts the windward side of the structure. Moreover, failures of structural components<br />

are interdependent. After the initial structural damage, a change in wind direction can<br />

significantly affect the subsequent occurrences of damage. This effect of interdependent<br />

component failures cannot be quantified reliably without considering the change in wind<br />

speed <strong>and</strong> wind direction over time within individual strong-wind events.<br />

Currently, two ways of specifying wind characteristics are commonly adopted in differ-<br />

ent vulnerability models. In the first method, used in damage estimation models such as<br />

HAZUS-MH (Vickery et al., 2006b), the structural loading <strong>and</strong> per<strong>for</strong>mance (of individual<br />

structures) are evaluated at a series of time steps <strong>for</strong> wind speed <strong>and</strong> direction; structural


Chapter 4. Integrated Multi-Structural Vulnerability Model 72<br />

damage at a particular time step depends on the cumulative damage in previous time steps.<br />

The second method, used in the FPHLP model (Gurley et al., 2006) <strong>and</strong> some other stud-<br />

ies (e.g. Li <strong>and</strong> Ellingwood, 2006), evaluates the structural per<strong>for</strong>mance at one given wind<br />

speed <strong>and</strong> direction. In some studies such as the FPHLP model, the average per<strong>for</strong>mance of<br />

the structure over all wind directions at a certain wind speed is then calculated. Progressive<br />

damage due to evolving wind speed/direction during a hurricane event is not accounted <strong>for</strong>.<br />

The first method gives a more detailed analysis of the effect of a storm passage on structure.<br />

The second method uses wind speed as the indicator of storm intensity, allowing easier com-<br />

putation <strong>and</strong> construction of vulnerability <strong>and</strong> fragility curves. Herein, we consider both<br />

analysis methods to assess damage to a cluster of buildings.<br />

4.2.2 Structural Characteristics<br />

This study analyzes hurricane wind damage to structures that are part of a residential<br />

community, as opposed to individual prototype structures (e.g. Gurley et al., 2006; Vickery<br />

et al., 2006b). A cluster of buildings is defined by specifying the location, orientation <strong>and</strong><br />

(significant) structural details of each building in a study area.<br />

A residential structure is composed of the structural system, enclosure system, mechani-<br />

cal <strong>and</strong> electrical system, interior system <strong>and</strong> its contents. The wind loads on the structural<br />

system <strong>and</strong> parts of the enclosure system can be calculated using structural engineering<br />

theories <strong>and</strong> by identifying their load paths. However, the loads on other systems, the<br />

interior system in particular, are technically infeasible to be determined accurately in engi-<br />

neering scale. There<strong>for</strong>e, the simulation engine cannot evaluate the hurricane damage on all<br />

systems, <strong>and</strong> the residential structure has to be simplified in the vulnerability model. The<br />

followings are building components on which the wind loadings can be obtained readily in<br />

engineering scale <strong>and</strong> are also significantly vulnerable to hurricane damage according to the<br />

investigation report of the FPHLP model (Gurley et al., 2005a): (1) roof covers, (2) roof


Chapter 4. Integrated Multi-Structural Vulnerability Model 73<br />

sheathing panels, (3) roof-to-wall connections, (4) gable-end sheathing panels (in gable roof<br />

only), (5) wall sheathing panels (in wood frame structures only), (6) walls, (7) openings<br />

(including windows, doors <strong>and</strong> garage door). A structural model is built by all the above<br />

components only.<br />

Currently considered building types include one-story concrete <strong>and</strong> wood-frame houses<br />

with gable roof or hip roof, of arbitrary overall dimensions, roof slope, sheathing panel size,<br />

truss spacing <strong>and</strong> number of openings. The two structure types above account <strong>for</strong> more<br />

than 70% of the residential structures in three of the four regions in Florida (Gurley et al.,<br />

2006). Statistics regarding house types <strong>and</strong> house geometries within a residential commu-<br />

nity may be obtained from in<strong>for</strong>mation about the local building stock (as in many cases<br />

building-specific data is not available). The statistics of the building components, which are<br />

inherently r<strong>and</strong>om in nature, may be found in many previous studies (e.g. National Asso-<br />

ciation of Home Builders, 1999; Rosowsky <strong>and</strong> Cheng, 1999b). Table 4.1 shows an example<br />

of the parameters of a typical residential structure’s geometry (Gurley et al., 2006).<br />

Table 4.1: Example parameters of a residential structure’s geometry.<br />

Parameter Value<br />

Length of structure 60 ft<br />

Width of structure 44 ft<br />

Height of wall 10 ft<br />

Roof pitch 5/12<br />

Eave overhang 2 ft<br />

Space between roof trusses 2 ft<br />

Roof sheathing panel dimension 8 ft × 4 ft<br />

Gable-end sheathing panel dimension 8 ft × 4 ft<br />

Wall sheathing panel 8 ft × 4 ft<br />

Figures 4.2 show the three dimensional renderings of four types of residential struc-<br />

ture models in this study. In the figure, the roof sheathing, gable-end sheathing <strong>and</strong> wall<br />

sheathing are lifted off in order to show the walls <strong>and</strong> the roof trusses clearly.


Chapter 4. Integrated Multi-Structural Vulnerability Model 74<br />

(a) (b)<br />

(c) (d)<br />

Figure 4.2: Three dimensional rendering of residential structure models: (a) concrete block<br />

house with gable roof (b) wood frame house with gable roof (c) concrete block house with<br />

hip roof (d) wood frame house with hip roof.


Chapter 4. Integrated Multi-Structural Vulnerability Model 75<br />

4.2.3 Component-Based Pressure Damage Model<br />

The component-based pressure damage model is developed based on the FPHLP model<br />

(Gurley et al., 2006). For a given wind speed <strong>and</strong> direction, wind loads on structural com-<br />

ponents are calculated based on the specified wind condition <strong>and</strong> designated load paths.<br />

All the building components in the models are assumed to fail only in their designated limit<br />

states listed in Table 4.2. All the failures except the projectile impact of openings may be<br />

determined in the component-based pressure damage model.<br />

Table 4.2: Limit states of structural components.<br />

Structural Component Limit State<br />

Roof Covers Uplift<br />

Roof Sheathing panels Uplift<br />

Roof-to-Wall Connections Uplift<br />

Gable-end Sheathing Panels Separation<br />

Wall Sheathing Panels Separation<br />

Concrete Walls Shear, Uplift, or Bending<br />

Wood Frame Walls Shear, Lateral Force or Bending<br />

Openings Pressure Failure or Projectile Impact<br />

Specification of wind pressure coefficient <strong>and</strong> wind pressure zone distribution <strong>for</strong> each<br />

building component are based on the FPHLP model (Gurley et al., 2006) <strong>and</strong> the ASCE<br />

7 St<strong>and</strong>ard (American Society of Civil Engineers, 2003), which is the st<strong>and</strong>ard used in the<br />

United States as a basis <strong>for</strong> determining structural loads. In the ASCE 7 wind specification,<br />

wind loads are considered in two different scales: (1) Main wind <strong>for</strong>ce resisting system<br />

(MWFRS), which is a group of structural members working together to resist <strong>and</strong> transfer<br />

wind loads acting on the entire structure to the ground; (2) Component <strong>and</strong> cladding (C&C),<br />

which is individual building component which receives wind load either directly or from the<br />

cladding, <strong>and</strong> transfers the loads to the MWFRS. In general, the C&C components are<br />

exposed to higher wind pressures than the MWFRS components. Given the configuration


Chapter 4. Integrated Multi-Structural Vulnerability Model 76<br />

of the structure model in this project, MWFRS coefficients are used only when wall shear<br />

<strong>for</strong>ce is calculated. <strong>Wind</strong> loadings on all other components adopt the C&C coefficients.<br />

Figure 4.3 illustrates the framework of loading mechanism using the ASCE 7 st<strong>and</strong>ard. In<br />

the modification of the ASCE 7 specification <strong>for</strong> hurricane loss estimation, all safety factors<br />

that are incorporated in the specification are removed <strong>for</strong> more accurate estimation of wind<br />

loading.<br />

Input<br />

Angle of<br />

<strong>Wind</strong><br />

Incidence<br />

3-second<br />

Gust <strong>Wind</strong><br />

Speed<br />

Sampled<br />

Resistance<br />

Capacities<br />

MWFRS<br />

C&C<br />

Wall Shear<br />

Wall Bending<br />

Moment / "<br />

Out-of Plane<br />

Roof Sheathing<br />

Gable-end<br />

Sheathing "<br />

(Gable Roof only)<br />

Wall Sheathing"<br />

(Wood Frame House<br />

only)<br />

Openings<br />

Roof Cover<br />

Wall Failure Check<br />

Wall Uplift<br />

Roof-to-Wall<br />

Connection<br />

Internal<br />

Pressure<br />

Output<br />

Area % of Failed<br />

Roof Cover<br />

Area % of Failed<br />

Roof Sheathing<br />

Area % of Failed<br />

Gable-end<br />

Sheathing "<br />

(Gable Roof Only)<br />

Area % of Failed<br />

Wall Sheathing (Wood<br />

Frame House Only)<br />

% of Failed"<br />

Roof-to-Wall<br />

Connections<br />

Number of<br />

Damaged Walls<br />

Number of Failed<br />

<strong>Wind</strong>ows<br />

Number of Failed<br />

Doors<br />

Number of Failed<br />

Garage Door<br />

Internal Pressure<br />

Figure 4.3: Flowchart of the component-based pressure damage model.<br />

By comparing the sample component resistance to the load, the pressure damage model<br />

assesses the wind-pressure damage to each building component. The levels of damage to<br />

these building components are highly correlated in most cases. When opening damage (<strong>for</strong>


Chapter 4. Integrated Multi-Structural Vulnerability Model 77<br />

instance, to windows, doors, or garage doors) occurs due to pressure or debris damage, the<br />

internal pressure is adjusted to a weighted-average external pressure acting on the areas<br />

of the broken components (see Gurley et al., 2006); this affects the damage conditions of<br />

almost all components. Different from the FPHLP model that adjusts the internal pressure<br />

once, in this integrated model the interplay between the internal pressure <strong>and</strong> the damage<br />

condition continues after the first internal-pressure adjustment. If additional openings are<br />

damaged due to the change in internal pressure, the internal pressure is updated again. An<br />

iterative algorithm is applied to obtain the equilibrium between the internal pressure <strong>and</strong><br />

the damage condition.<br />

4.2.4 <strong>Wind</strong>borne Debris Model<br />

In tropical cyclones, windborne debris significantly contributes to the damage of building en-<br />

velope, particularly glass openings (National Institute of St<strong>and</strong>ards <strong>and</strong> Technology, 2006b).<br />

For simplicity, we assume that glass windows are the only structural members vulnerable<br />

to windborne debris. Both HAZUS-MH hurricane model (Federal Emergency <strong>Management</strong><br />

Agency, 2006) <strong>and</strong> the FPHLP Model (Gurley et al., 2006) <strong>for</strong>mulate the probability of<br />

debris damage to an opening, PV (D), at a given wind speed (V ) as:<br />

PV (D) = 1 − exp[−λq(1 − P (ζ < ζ0))], (4.1)<br />

where λ is the mean number of missile impacts on the building, q is the fraction of the<br />

building surface covered by the opening, P (ζ − ζ0) is the probability that the momentum<br />

(ζ) of the debris is less than the damage threshold value (ζ0) given an impact. In the<br />

FPHLP Model (Gurley et al., 2006), the parameters λ <strong>and</strong> P (ζ − ζ0) are some functions of<br />

wind speed (V ) only <strong>and</strong> are independent of all other factors.<br />

Alternatively, λ <strong>and</strong> P (ζ − ζ0) may be simulated based on the potential number of<br />

debris objects in the environment <strong>and</strong> the properties of the debris objects. We assume that


Chapter 4. Integrated Multi-Structural Vulnerability Model 78<br />

the only debris sources are the roof covering <strong>and</strong> sheathing panel uplifted by wind pressure<br />

from nearby residences. Post-damage survey by Twisdale et al. (1996) shows that the debris<br />

either flies less than 1 second (mean = 0.73 sec), or stays in the air <strong>for</strong> a much longer time<br />

(mean = 1.58 sec). The parameters λ <strong>and</strong> P (ζ − ζ0) are calculated independently <strong>for</strong> each<br />

flight mode. Their weighted average will be used <strong>for</strong> equation (4.1) to estimate PV (D).<br />

Properties of potential windborne debris objects, such as the mass <strong>and</strong> area of roof cover<br />

<strong>and</strong> roof sheathing panels, plus the glass windows resistance against debris impact, are re-<br />

quired to quantify the windborne debris risk. The properties of common debris objects are<br />

presented in several studies (e.g. McDonald, 1990; Minor et al., 1978). Table 4.3 lists the<br />

properties of debris objects from the study by Twisdale et al. (1996).<br />

Table 4.3: Properties of debris objects.<br />

Parameter Symbol Description Value<br />

Flight Time td Mode 1 (P =0.38) 0.73 sec<br />

Mode 2 (P =0.62) 1.58 sec<br />

Debris Mass md Roof cover 2.5 lbf<br />

Sheathing panel 32 lbf<br />

Debris Area Ad Roof cover 3 ft 2<br />

Sheathing panel 32 ft 2<br />

The mean along-wind distance traveled by a debris object, ¯ Xd , is approximated by an<br />

empirical equation (Lin et al., 2006):<br />

¯Xd = 2Md<br />

[<br />

Adρa<br />

1<br />

2 C(Kt′ ) 2 + c1(Kt ′ ) 3 + c2(Kt ′ ) 4 + c3(Kt ′ ) 5 ], (4.2)<br />

where C is a debris geometry constant; c1, c2 <strong>and</strong> c3 = regression coefficients; t ′ = gt/V ;<br />

<strong>and</strong> K is the Tachikawa number (Tachikawa, 1983) given by:<br />

K = ρaV 2Ad , (4.3)<br />

2mdg


Chapter 4. Integrated Multi-Structural Vulnerability Model 79<br />

where Ad is debris area; md is debris mass; <strong>and</strong> ρa is air density. Equation (4.3) applies<br />

when Kt ′ is between [0, 6.5]. If Kt ′ > 6.5, ¯ Xd is the value approximated by using Kt ′ =<br />

6.5, plus an additional distance by assuming that the flight speed is equal to 0.98V . The<br />

along-wind displacement Xd is normally distributed with a COV of 0.35. The across-wind<br />

distance Yd is normally distributed with a mean of 0 <strong>and</strong> a COV of 0.35. It is assumed that<br />

the debris hits the structure if it falls on the plan area of the target structure.<br />

The debris impact velocity, Vd, is assumed to follow a beta distribution (Lin et al. 2009),<br />

with probability density function:<br />

d (1 − vd) b−1<br />

va−1<br />

f(vd)vd =<br />

B(a, b)<br />

, (4.4)<br />

where B(·) denotes a Beta function. The parameters a <strong>and</strong> b are determined by assuming<br />

the mean of debris velocity <strong>and</strong> the dispersion parameter are:<br />

¯Vd = a<br />

<br />

CρaAd¯x<br />

= V 1 − exp −<br />

, (4.5)<br />

a + b<br />

<br />

1<br />

η = a + b = max ,<br />

¯Vd<br />

1<br />

1 − ¯ Vd<br />

Md<br />

<br />

+ γ, (4.6)<br />

where γ is assumed to be 3 (Lin et al. 2009), which may be statistically updated when<br />

new debris data is available. The probability that the momentum, ζ = mdVd, exceeds the<br />

damage threshold is hence obtained.<br />

4.2.5 Integrated Structural Damage Estimation<br />

The Lin et al. (2010) vulnerability model integrates the debris risk model <strong>and</strong> the component-<br />

based pressure damage model. Damaged roof cover, roof sheathing, <strong>and</strong> gable-end sheathing<br />

are considered as debris sources, <strong>and</strong> windows <strong>and</strong> glass doors (or their protective covers) are<br />

assumed to be debris-impact vulnerable areas. The integrated structural damage estimation


Chapter 4. Integrated Multi-Structural Vulnerability Model 80<br />

INPUT PRESSURE DAMAGE MODEL DEBRIS DAMAGE MODEL<br />

OUTPUT<br />

Structure 1 of N<br />

Structure 1 of N<br />

Y<br />

N<br />

New<br />

Opening<br />

Damage<br />

Debris<br />

Damage<br />

Analysis<br />

Internal<br />

Pressure<br />

Adjustment<br />

Y<br />

New<br />

Opening<br />

Damage<br />

Pressure<br />

Damage<br />

Analysis<br />

Any Roof<br />

Cover or<br />

Sheathing<br />

Damage<br />

from Any<br />

Structure<br />

=<br />

Available<br />

Debris<br />

Objects<br />

<strong>Wind</strong><br />

Speed<br />

N<br />

Final<br />

Damage<br />

Condition<br />

of<br />

<strong>Structures</strong><br />

1 - N<br />

Structure 2 of N<br />

Structure 2 of N<br />

<strong>Wind</strong><br />

Direction<br />

...<br />

...<br />

Structure N-1 of N<br />

Structure N-1 of N<br />

Structure N of N<br />

Structure N of N<br />

Y<br />

Initial<br />

Damage<br />

Condition<br />

of<br />

<strong>Structures</strong><br />

1 - N<br />

N<br />

New<br />

Opening<br />

Damage<br />

Debris<br />

Damage<br />

Analysis<br />

Internal<br />

Pressure<br />

Adjustment<br />

Y<br />

New<br />

Opening<br />

Damage<br />

Pressure<br />

Damage<br />

Analysis<br />

N<br />

If it is a time series analysis, the final damage condition in each structure<br />

becomes the initial damage condition in the next step.<br />

Figure 4.4: Flowchart of the integrated vulnerability model.


Chapter 4. Integrated Multi-Structural Vulnerability Model 81<br />

model is summarized in a flowchart in Figure 4.4. The model implements Monte Carlo<br />

simulation, with structural resistance r<strong>and</strong>omly assigned to every building component in<br />

each simulation run, in accordance with the specified (joint) probability distributions. In<br />

most cases, the structural parameters are assumed independent of each other. However,<br />

the roof-to-wall connection strength in each structure is batch-selected to simulate the<br />

consistency in quality of material <strong>and</strong> installation within a single structure (see Gurley<br />

et al., 2006).<br />

In each simulation, <strong>for</strong> a given wind speed <strong>and</strong> direction, wind loads <strong>and</strong> damages<br />

of structural components are first calculated using the component-based pressure damage<br />

model. After that, the in<strong>for</strong>mation about roof covers <strong>and</strong> roof sheathing panels damaged by<br />

wind pressure is passed along to the debris risk model, which estimates the probability of<br />

debris damage to each window <strong>and</strong> glass door of each house. If there is any debris damage,<br />

the internal pressure of the damaged house is updated, <strong>and</strong> the damage to structural com-<br />

ponents is recalculated in the pressure damage model. The in<strong>for</strong>mation about any newly<br />

damaged roof covers <strong>and</strong> roof-sheathing panels are again sent to the debris risk model to<br />

estimate their damage risk, starting a loop of pressure-debris damage interaction, until no<br />

more debris is generated from any house. The damage condition of each building is thus<br />

obtained <strong>for</strong> the given wind speed <strong>and</strong> direction. In the case of time-series damage analysis,<br />

similar analysis is conducted <strong>for</strong> the wind speed <strong>and</strong> direction at each analysis step of the<br />

wind time history; the damage condition of each structure from one analysis step is carried<br />

to the next step, until the end of the hurricane wind time series.<br />

4.2.6 Approximate Economic Model<br />

Based on the percentage damage to each structural component, the damage ratio of a<br />

building is approximated, following (e.g. Leicester et al., 1979), as a ratio of total repair<br />

cost to the initial (or replacement) cost of the structural system. To calculate the damage


Chapter 4. Integrated Multi-Structural Vulnerability Model 82<br />

ratio, the breakdown of the initial cost of the building is obtained from construction cost<br />

databases such as R.S. Means Residential Cost Data (R.S. Means., 2009). The damage<br />

ratio of the structural system is the weighted average of the damage percentages of all<br />

structural components considered in the integrated vulnerability model (roof cover, roof<br />

sheathing panels, etc.), with the weights proportional to the repair cost of the corresponding<br />

component.<br />

Similarly, damage ratios may be defined <strong>for</strong> the interior system <strong>and</strong> the utility system,<br />

which, combined with the structural system, constitute an entire residential structure. How-<br />

ever, unlike <strong>for</strong> the structural system, the damages on interior system <strong>and</strong> utility system are<br />

not readily analyzable using (structural) engineering models. There<strong>for</strong>e, loss models such<br />

as HAZUS-MH (Vickery et al. 2006) <strong>and</strong> FPHLP (Gurley et al., 2006) <strong>for</strong>mulate sets of<br />

empirical equations to relate the interior damage to the percentage damage of each struc-<br />

tural component. Similar equations are also applied to modeling utility-system damage.<br />

The total damage ratio of a building, which is a measure of insurance loss, is the weighted-<br />

average of the damage ratios of each subassembly, with the weights proportional to the<br />

respective repair costs. (Content loss, additional living expenses, <strong>and</strong> business interruption<br />

are usually also counted in economic loss estimation in the insurance industry. They are not<br />

included in this study, however, since they greatly depend on building function, occupancy,<br />

<strong>and</strong> details of insurance policies.) Since the pressure damage module of the integrated vul-<br />

nerability model is based on the FPHLP model, <strong>for</strong> consistency, this study employs the<br />

FPHLP method to evaluate the damage ratios <strong>for</strong> the interior system <strong>and</strong> utility system.


Chapter 4. Integrated Multi-Structural Vulnerability Model 83<br />

4.3 Numerical Examples<br />

4.3.1 A Hip-roof Concrete Residential Structure in Isolation<br />

Be<strong>for</strong>e showing the results of the integrated vulnerability model, a hip-roof concrete house<br />

is used as an example to illustrate the component-based pressure damage model. In this<br />

example, there is no windborne debris impact on the structure.<br />

Figure 4.5 illustrates the vulnerability of a hip-roof concrete house at four different wind<br />

speeds from 100 mph to 250 mph. 1,000 simulations were carried out at each of the eight<br />

nominal wind directions at each wind speed. The results are the average vulnerability of<br />

the house over the eight wind directions. In the figure, the color indicates the probability<br />

of failure of each structural component. As wind speed increases, the probability of failure<br />

increases <strong>for</strong> every component. Virtually all the roof covers are damage when wind speed<br />

reaches 250 mph. If this is combined with the windborne debris model, the failed roof<br />

panels will serve as a considerable amount of debris objects that impact the other nearby<br />

structures. Figure 4.6 illustrates the vulnerability of the same hip-roof concrete house under<br />

a 200-mph wind in three different directions.<br />

4.3.2 Hypothetical Residential Community<br />

This second numerical example is to illustrate the effect of windborne debris on the neigh-<br />

boring structures. Consider a residential development of 16 identical single-story concrete<br />

block gable roof houses having the same characteristics <strong>and</strong> orientation, as shown in Figure<br />

4.7. Vulnerability analysis of the hypothetical houses is conducted <strong>for</strong> a wind speed of 65<br />

m/s at 45-degree angle, based on 500 Monte Carlo simulations. Since all the houses in this<br />

hypothetical development are identical, the difference in damage to each house is mainly<br />

caused by the debris impact <strong>and</strong> the subsequent internal pressure change.<br />

Figure 4.8 presents the breakdown of estimated structural damage to each house. Note


Chapter 4. Integrated Multi-Structural Vulnerability Model 84<br />

(a) 100 mph (b) 150 mph<br />

(c) 200 mph (d) 250 mph<br />

Figure 4.5: Vulnerability of a hip-roof concrete house at four different wind speeds.


Chapter 4. Integrated Multi-Structural Vulnerability Model 85<br />

(a) 0 ◦ wind angle (b) 45 ◦ wind angle<br />

(c) 90 ◦ wind angle<br />

Figure 4.6: Vulnerability of a hip-roof concrete house under 200-mph at three different wind<br />

directions.


Chapter 4. Integrated Multi-Structural Vulnerability Model 86<br />

Figure 4.7: Three-dimensional view of a hypothetical residential development. The arrows<br />

indicate wind direction.<br />

that the house at the upper right corner is hardly impacted by debris, while all houses at<br />

the lower left are affected by debris impact. There is a trend indicating that the further the<br />

house is from the upper right, the more damage it suffers. Among all structural components,<br />

the window <strong>and</strong> door are the two most damaged components, as both are debris-vulnerable<br />

according to the model. The damage to other structural components on suction zones, such<br />

as garage doors, roof sheathings, <strong>and</strong> roof covers, are also increased due to the internal<br />

pressurization resulting from debris damage.<br />

The in<strong>for</strong>mation about structural damage presented in Figure 4.8 is converted to sub-<br />

assembly damage ratios <strong>and</strong> an overall damage ratio. The average damage ratios of all<br />

houses are shown in Figure 4.9 as bars labeled debris. The vulnerability analysis is then<br />

repeated <strong>for</strong> the same residential development under the same wind condition, but without<br />

accounting <strong>for</strong> any debris effects. The resulting damage ratios are also shown in Figure<br />

4.9, as bars labeled no debris. Notice significant differences between the damage ratios from<br />

the two analysis methods. If debris is not considered, the average overall damage ratio is


LEGEND<br />

Roof sheathing<br />

onnection<br />

indow<br />

Door<br />

Connection<br />

1% Damage<br />

indow<br />

Damage<br />

Garage door<br />

Roof cover<br />

12.5%<br />

25%<br />

37.5% Wall<br />

50%<br />

EXAMPLE<br />

Roof sheathing<br />

7% Damage<br />

Door<br />

3% Damage Garage door<br />

15% Damage<br />

Roof cover<br />

21% Damage<br />

Wall<br />

1% Damage<br />

Chapter 4. Integrated Multi-Structural Vulnerability Model 87<br />

Longitude (m)<br />

90<br />

60<br />

30<br />

0<br />

DAMAGE ESTIMATON OF HYPOTHETICAL RESIDENTIAL DEVELOPMENT<br />

LEGEND<br />

Roof sheathing<br />

Connection<br />

Roof cover<br />

<strong>Wind</strong>ow<br />

0 30 60<br />

Latitude (m)<br />

12.5%<br />

25%<br />

37.5% Wall<br />

60<br />

DAMAGE ESTIMATON 50% OF HYPOTHETICAL RESIDENTIAL DE<br />

Door 90<br />

Garage door<br />

LEGEND<br />

Roof sheathing<br />

Connection<br />

<strong>Wind</strong>ow<br />

Door<br />

Roof cover<br />

12.5%<br />

25%<br />

37.5% Wall<br />

50%<br />

Garage door<br />

EXAMPLE<br />

Longitude (m)<br />

90<br />

60<br />

0<br />

Connection<br />

11% Damage<br />

<strong>Wind</strong>ow<br />

41% Damage<br />

EXAMPLE<br />

Roof sheathing<br />

7% Damage<br />

Door<br />

43% Damage Garage door<br />

15% Damage<br />

Roof cover<br />

21% Damage<br />

Wall<br />

1% Damage<br />

Figure 4.8: Estimated damage condition of each individual house in the hypothetical residential<br />

development under a wind speed of 65 m/s at a 45-degree angle. Each radar<br />

30<br />

Roof sheathing<br />

diagram represents the mean damage percentage of the structural components of an indi-<br />

7% Damage<br />

Connection<br />

vidual house. The 11% arrows Damage indicate wind direction. On the left panel are the legend <strong>and</strong> an<br />

Roof cover<br />

example illustrating the meaning21% ofDamage the radar diagram.<br />

<strong>Wind</strong>ow<br />

41% Damage<br />

Door<br />

43% Damage Garage door<br />

15% Damage<br />

Wall<br />

1% Damage<br />

Longitude (m)<br />

90<br />

30<br />

0<br />

DAMAGE ES<br />

0 30 60 9<br />

0


Chapter 4. Integrated Multi-Structural Vulnerability Model 88<br />

Damage Ratio<br />

0.16<br />

0.14<br />

0.12<br />

0.1<br />

0.08<br />

0.06<br />

0.04<br />

0.02<br />

Debris<br />

No debris<br />

0<br />

Structural System Interior Utility System Overall<br />

Figure 4.9: Comparison of subassembly <strong>and</strong> overall damage ratios with <strong>and</strong> without consideration<br />

of windborne debris. The bars labeled debris indicate the average damage ratios<br />

of all houses in the hypothetical residential development if windborne debris is considered.<br />

The bars labeled no debris indicate the same variables without considering debris.<br />

underestimated, in this case, by as much as 29%.<br />

4.3.3 Two Neighboring Houses<br />

The previous numerical example demonstrated the effect of neighboring buildings on a<br />

structures wind damage on account of the locations of the buildings. This second example<br />

further illustrates how a deteriorated neighboring building may increase damage through<br />

windborne debris. Consider two houses separated by 30 m (see Figure 4.10a). Both are<br />

the same as the houses in the previous example, except that the resistance of all building<br />

components in House B is multiplied by a resistance factor.<br />

Figure 4.10b shows the overall damage ratio (<strong>and</strong> its breakdown) <strong>for</strong> structure A versus<br />

the resistance factor of structure B. While structure As resistance is always the same, an<br />

increase in structure Bs resistance lowers the windborne debris risk faced by structure A,<br />

resulting in a lower damage ratio of structure A. In this particular example, if the resistance<br />

of a neighboring house deteriorates <strong>and</strong> becomes half of that of another house, it can cause


Chapter 4. Integrated Multi-Structural Vulnerability Model 89<br />

Structure A<br />

Structure B<br />

30m<br />

<strong>Wind</strong> Direction<br />

Damage Ratio of Structure A<br />

0.35<br />

0.3<br />

0.25<br />

0.2<br />

0.15<br />

0.1<br />

0.05<br />

Structural System<br />

Utility System<br />

Interior System<br />

0<br />

0.5 1<br />

Resistance Factor of Structure B<br />

1.5<br />

(a) (b)<br />

Figure 4.10: (a) Illustration of two neighboring houses. (b) Breakdown of overall damage<br />

ratio of structure A versus resistance factor of structure B.<br />

about 90% more damage to the other house.<br />

4.3.4 Sarasota, Florida<br />

This last numerical example demonstrates the difference in damage estimation results be-<br />

tween time series analysis <strong>and</strong> point-in-time analysis during a storm passage. Consider a<br />

residential development of 358 single-story houses in Sarasota County, Florida (Figure 4.11).<br />

The locations of the residential houses are known, while the orientation of each house is<br />

modeled as a Gaussian r<strong>and</strong>om variable with as its mean the direction perpendicular to the<br />

main street of the house <strong>and</strong> as its st<strong>and</strong>ard deviation 15 degrees. The structural character-<br />

istics of each house are estimated based on in<strong>for</strong>mation about the building stock in central<br />

Florida (Gurley et al., 2006). We estimate wind damage to the study area from Hurricane<br />

Charley (2004), which passed to the left of the study area. Hurricane Charley is numerically<br />

simulated using the Weather Research <strong>and</strong> Forecasting (WRF) model (Skamarock et al.,<br />

2005); this yields a simulated wind-velocity time history at a representative location within<br />

the study region.


Chapter 4. Integrated Multi-Structural Vulnerability Model 90<br />

Figure 4.11: Study area of a residential development of 358 houses in Sarasota County,<br />

Florida (left) <strong>and</strong> simulated storm track of Hurricane Charley of 2004 (right). The circle<br />

(in both panels) represents the recording location of the wind time history during the<br />

passage of the simulated storm.<br />

Unlike most other vulnerability models (Gurley et al., 2006; Vickery et al., 2006b) that<br />

evaluate individual prototype buildings, the integrated vulnerability model analyzes a clus-<br />

ter of buildings as a whole, seeking to account <strong>for</strong> the (time-varying) interaction between<br />

building damage levels due to windborne debris. In this case of a 358-residence community,<br />

computation time <strong>for</strong> estimating the damage condition can be significant. Evaluation of<br />

building per<strong>for</strong>mance only at the maximum wind speed (<strong>and</strong> the corresponding direction),<br />

has the advantage, compared to the cumulative damage analysis, of reduced computational<br />

ef<strong>for</strong>t. The results of both methods of representing the wind conditions are compared in<br />

Figure 4.12.


DAMAGE ANALYIS<br />

USING MAXIMUM WIND VELOCITY<br />

Chapter 4. Integrated Multi-Structural Vulnerability Model 91<br />

DAMAGE ANALYSIS USING A TIME SERIES OF WIND VELOCITY<br />

Only Step<br />

Step 1 Step 2 Step 3 Step 4 Step 5 Step 6 Step 7 Step 8<br />

150<br />

100<br />

50<br />

<strong>Wind</strong> speed<br />

(mph)<br />

0<br />

360 5 10 15 20<br />

240<br />

120<br />

0<br />

0 5 10 15 20<br />

Time (hour)<br />

<strong>Wind</strong> direction<br />

(deg)<br />

Roof cover<br />

2.4%<br />

150<br />

100<br />

50<br />

0<br />

360<br />

5 10 15 20<br />

240<br />

120<br />

0<br />

0 5 10 15 20<br />

Time (hour)<br />

Roof sheathing<br />

Roof sheathing<br />

LEGEND<br />

EXAMPLE Connection 0.2%<br />

Connection<br />

0.0% damage<br />

Roof cover<br />

<strong>Wind</strong>ow<br />

3.7% damage<br />

<strong>Wind</strong>ow<br />

Wall<br />

0.0%<br />

Wall<br />

1.25%<br />

2.5%<br />

3.75%<br />

5%<br />

Door<br />

3.3% damage Garage door<br />

0.1% damage<br />

Door Garage door<br />

Figure 4.12: Comparison of the estimated wind damage condition of the residential development in Sarasota County,<br />

Florida, under a Hurricane Charley wind condition, by time series (left) <strong>and</strong> maximum wind analyses (right), respectively.<br />

The radar diagrams represent the overall average damage percentage of the building components of all houses, under the<br />

simulated wind time history of Hurricane Charley of 2004. The red dots indicate the analysis wind series of the maximum<br />

wind velocity in each 15-degree direction segment. The legend with an example illustrates the meaning of the radar<br />

diagram.


Chapter 4. Integrated Multi-Structural Vulnerability Model 92<br />

In the time-series damage analysis, structural damage is estimated over a time series of<br />

wind speed <strong>and</strong> wind direction. Figure 4.12 (left) shows the analysis result using a sequence<br />

of maximum wind speeds (<strong>and</strong> corresponding wind directions) during consecutive 15-degree<br />

wind-direction segments of the wind time history. Note that the analysis steps are unevenly<br />

distributed, being more concentrated where the wind direction varies significantly. Analysis<br />

using a 15-min interval time series was also conducted, but the results are not shown because<br />

they are very similar to those based on the 15-degree wind-direction segmentation (<strong>for</strong> which<br />

computational ef<strong>for</strong>t is considerably reduced (see Lin et al., 2010). As Figure 4.12 (left)<br />

shows, significant damage starts to occur, in this case, when the wind speed approaches its<br />

peak. Due to the relatively low wind speeds (compared to 65 m/s in the previous numerical<br />

example), only the most vulnerable building components, including roof covers, windows<br />

<strong>and</strong> doors, sustain damage.<br />

Damage Ratio<br />

0.01<br />

0.009<br />

0.008<br />

0.007<br />

0.006<br />

0.005<br />

0.004<br />

0.003<br />

0.002<br />

0.001<br />

Progressive<br />

Max <strong>Wind</strong><br />

0<br />

Structural System Interior Utility System Overall<br />

Figure 4.13: Comparison of subassembly <strong>and</strong> overall damage ratios between time-series<br />

<strong>and</strong> maximum-wind damage analyses. The bars labeled as progressive indicate the average<br />

damage ratios of all houses in Sarasota County, Florida using the analysis wind series of<br />

the maximum wind velocity in each 15-degree direction segment of the wind time history.<br />

The bars labeled as max wind indicate the same variables using only the maximum wind<br />

speed <strong>and</strong> the corresponding wind direction.<br />

Alternatively, the damage condition is estimated using the maximum wind speed (<strong>and</strong>


Chapter 4. Integrated Multi-Structural Vulnerability Model 93<br />

corresponding direction) <strong>for</strong> the entire wind time history. Computationally faster since it<br />

involves a one-step analysis compared to multiple steps in the time-series damage analysis,<br />

this method predicts significantly lower damage. As shown in Figure 4.13, the predicted<br />

overall average damage ratio can differ by about 25% <strong>for</strong> the different methods.<br />

4.4 Closing Remarks<br />

We presented <strong>and</strong> illustrated improved methodology to assess structural vulnerability <strong>and</strong><br />

wind-related economic losses in residential neighborhoods during hurricanes. Involving the<br />

interplay between a pressure damage model <strong>and</strong> a debris risk model, the method is capable of<br />

estimating cumulative damage to the overall structural system, the interior, <strong>and</strong> the utility<br />

system of buildings in a residential development. By means of four numerical examples, it is<br />

shown that windborne debris tends to contribute significantly to total damage <strong>and</strong> economic<br />

loss, <strong>and</strong> that the cumulative wind-related damage during the passage of a storm may be<br />

much greater than would be estimated by considering only the maximum wind speed during<br />

the storm. Future field studies are needed to collect damage data to validate this (<strong>and</strong> other)<br />

damage assessment methodology. The integrated vulnerability model will be applied to<br />

develop vulnerability <strong>and</strong> fragility curves <strong>for</strong> building categories within generic communities<br />

in hurricane-prone regions. To further improve wind damage estimation <strong>for</strong> residential<br />

communities, in addition to debris effects, more complex <strong>and</strong> detailed aerodynamic effects<br />

may have to be incorporated into the models.


Chapter<br />

5<br />

Long-Term <strong>Risk</strong> <strong>Assessment</strong> <strong>and</strong><br />

<strong>Management</strong><br />

5.1 Introduction<br />

Effects of climate change have been extensively studied in various scientific fields. How-<br />

ever, the potential impact of climate change on structural engineering analysis has been<br />

addressed by very limited literature only (Kasperski, 1998; Steenbergen et al., 2009). In<br />

current practice of civil engineering, environmental loads are usually calculated based on<br />

data collected over the past decades, assuming that they have the same trend in the future.<br />

This omission of climate change may underestimate the future structural loads, which may<br />

particularly concern new infrastructure projects which are expected to service <strong>for</strong> more than<br />

50-100 years as well as existing buildings which will continue to be in use <strong>for</strong> decades.<br />

Buildings are impacted by climate change in many <strong>for</strong>ms including wind load, rain<br />

load, snow load <strong>and</strong> temperature load (Steenbergen et al., 2009). This study focuses on<br />

extreme wind load, which is one of the most common types of load causing structural<br />

failures. Given that the trends of future wind hazards are still uncertain, we present a non-<br />

stationary compound Poisson process to simulate future wind conditions under the effect<br />

94


Chapter 5. Long-Term <strong>Risk</strong> <strong>Assessment</strong> <strong>and</strong> <strong>Management</strong> 95<br />

of climate change. The simulated wind conditions are applied to two different vulnerability<br />

models: a single-limit-state vulnerability model <strong>and</strong> the modified Florida Public Hurricane<br />

Loss Projection (FPHLP) model, to illustrate the effect of climate change on a structure’s<br />

life-cycle cost <strong>and</strong> the decisions regarding the structural design.<br />

5.2 Methodology<br />

In the insurance literature, stochastic processes (e.g. Poisson process) are commonly em-<br />

ployed to simulate economic damage associated with extreme weather as a function of<br />

time (Embrechts et al., 1997; Katz, 2002). In order to more accurately assess the poten-<br />

tial effect of climate change, the simulation process is divided into two parts. First, a<br />

stochastic process, which is referred as the hazard model, is used to simulate the wind<br />

hazard events. Parameters describing the climate change scenarios are implemented in this<br />

stochastic process. Then, the simulated wind conditions are translated into the damage<br />

level (<strong>and</strong> associated economic cost) of an individual structure using vulnerability models<br />

that are developed using reliability engineering theories. Combining the hazard model <strong>and</strong><br />

the vulnerability model, of which the details are described in the following sections, we can<br />

obtain projections of future economic cost of a structure due to wind hazard in different<br />

climate change scenarios.<br />

Given the projections of future economic cost, the change in risk <strong>and</strong> reliability as a<br />

result of climate change may be quantified by a number of metrics. Life-cycle cost, which<br />

is the overall discounted cost in a structure’s design life, serves as an indicator of the<br />

cost-effectiveness of a structure <strong>and</strong> governs the objective function in the optimization of<br />

structural design (Frangopol <strong>and</strong> Maute, 2003; Chang <strong>and</strong> Shinozuka, 1996). Let [0, tL] be a<br />

structure’s life cycle <strong>and</strong> N(tL) be the total number of wind hazard events occurred within<br />

[0, tL]. Let Ti denote the time that the i-th wind event occurs. A structure’s total life-cycle


Chapter 5. Long-Term <strong>Risk</strong> <strong>Assessment</strong> <strong>and</strong> <strong>Management</strong> 96<br />

cost, C, may be expressed as:<br />

N(tL) <br />

C = C0 + δ(Ti)Cr,i, (5.1)<br />

i=1<br />

where C0 is the initial construction cost; Cr,i the repair cost associated with the i-th hazard<br />

event. The discount factor, δ(t), is used to to convert the future Cr,i into the present value.<br />

This study adopts the most common exponential discounting:<br />

δ(t) = exp(−γt), (5.2)<br />

where γ is the discount rate. Typical discount rates range from 5% to 20% (e.g. Coffelt<br />

et al., 2010), <strong>and</strong> a 5% discount rate will be used throughout this study. We are going<br />

to investigate the change in expected value <strong>and</strong> probability distribution of life-cycle cost,<br />

particularly the total discounted repair cost, in different scenarios of wind event frequency<br />

<strong>and</strong> intensity.<br />

5.3 <strong>Hazard</strong> model<br />

5.3.1 <strong>Wind</strong> Speed Distribution<br />

Severe winds, which originate from normal weather or different types of wind hazards (i.e.<br />

thunderstorms, tropical cyclones, tornadoes, etc.), may be characterized by wind speed,<br />

wind direction, duration, etc. <strong>Wind</strong> speed is conventionally used as the metric to measure<br />

a wind hazard’s intensity. For example, the Saffir-Simpson hurricane scale is based on<br />

maximum sustained wind speed. This study will also use wind speed as the parameter<br />

linking the hazard model <strong>and</strong> the vulnerability model.<br />

<strong>Wind</strong> speed is the governing factor of a structure’s resistance against lateral <strong>and</strong> uplift<br />

<strong>for</strong>ces in structural design. In the United States, the design wind speed is specified by the


Chapter 5. Long-Term <strong>Risk</strong> <strong>Assessment</strong> <strong>and</strong> <strong>Management</strong> 97<br />

50-year return period, which means the wind speed with a probability of being exceeded<br />

per year of 1 in 50. Probability distribution of wind speed may be obtained by best fitting<br />

observed data. While strong frontal depressions <strong>and</strong> thunderstorms may be adequately<br />

captured in the collected data over the past decades, the tropical cyclones may not be so<br />

due to their infrequent occurrence. Monte Carlo simulations are often employed to decide<br />

the design wind speed in hurricane-prone regions.<br />

Generalized extreme value (GEV) distribution is generally adopted to represent the<br />

annual maximum wind, Vmax:<br />

F (vmax) = exp<br />

<br />

−<br />

<br />

1 + ξ<br />

<br />

−1/ξ<br />

vmax − ψ<br />

, (5.3)<br />

σ<br />

where F (·) is the cumulative distribution function, <strong>and</strong> ψ, σ <strong>and</strong> ξ are the location, scale<br />

<strong>and</strong> shape parameters, respectively. If ξ → 0, the distribution is referred as Gumbel distri-<br />

bution (type I extreme value distribution), <strong>and</strong> if ξ < 0, it is refered as Weibull distribution<br />

(type III extreme value distribution). (ξ > 0 defines the Fréchet distribution (type II ex-<br />

treme value distribution), but it is not discussed here.) In meteorology <strong>and</strong> wind energy<br />

analysis, Weibull distribution is commonly used as the extreme wind speed distribution,<br />

since it has an upper limit, which is appropriate <strong>for</strong> physical grounds (Simiu <strong>and</strong> Heckert,<br />

1996). The use of Gumbel distribution <strong>for</strong> extreme wind distribution has a long history <strong>and</strong><br />

is still being extensively used (Galambos <strong>and</strong> Macri, 1999; Holmes, 2007; Naess <strong>and</strong> Gaidai,<br />

2009). Though it does not impose any upper limit on the wind speed like Weibull distri-<br />

bution, it is suitable <strong>for</strong> the purpose of engineering design codes <strong>and</strong> st<strong>and</strong>ards (Holmes,<br />

2007). In addition, Gumbel distribution describes the distribution of the maxima of a series<br />

of identical <strong>and</strong> independently distributed exponential r<strong>and</strong>om variables. This allows us<br />

to interpret the distribution as a compound Poisson process, which is useful in this risk<br />

assessment study because it allows simulation of the number of failure events as well as the<br />

time of each occurrence of failure event.


Chapter 5. Long-Term <strong>Risk</strong> <strong>Assessment</strong> <strong>and</strong> <strong>Management</strong> 98<br />

5.3.2 Compound Poisson Process<br />

Assume that the occurrence of wind event is a Poisson process with rate of occurrence λ.<br />

The probability mass function of the number of wind hazards, n, occurring in the time<br />

interval [0, t) is:<br />

P(N = n) = e−λt (λt) n<br />

, n ≥ 0. (5.4)<br />

n!<br />

Furthermore, let the intensity (wind speed) of each wind event (V1, . . . , Vn) be an identical<br />

exponential r<strong>and</strong>om variable which is independent of each other <strong>and</strong> has a mean of µ:<br />

F (vn) = 1 − e −vn/µ . (5.5)<br />

Based on the assumptions on the frequency <strong>and</strong> intensity of wind hazards, the cumulative<br />

distribution function of the maximum wind speed within the time interval [0, t), which is<br />

defined as Vmax = max(V1, . . . , Vn), is equal to:<br />

<br />

F (vmax) = exp −λt exp − vmax<br />

<br />

. (5.6)<br />

µ<br />

This is equivalent to the Gumbel cumulative distribution function, which is usually ex-<br />

pressed as:<br />

<br />

F (vmax) = exp − exp<br />

α − vmax<br />

β<br />

<br />

, (5.7)<br />

where α = µ ln(λt) is the location parameter <strong>and</strong> β = µ is the scale parameter. The return<br />

period, which is the inverse of the non-exceedence probability, is directly related to the<br />

cumulative probability distribution F (vmax) as follows:<br />

Return Period =<br />

1<br />

. (5.8)<br />

1 − F (vmax)<br />

The above stationary compound Poisson process may be converted to a non-stationary


Chapter 5. Long-Term <strong>Risk</strong> <strong>Assessment</strong> <strong>and</strong> <strong>Management</strong> 99<br />

process by introducing time-varying parameters λ(t) <strong>and</strong> µ(t) to replace λ <strong>and</strong> µ in equa-<br />

tions (5.4) <strong>and</strong> (5.5). This is similar to the non-stationary analysis of extreme wind speed<br />

distribution (e.g. Hundecha et al., 2008), in which the distribution parameters are fit to<br />

some functions of time, except that both parameters λ <strong>and</strong> µ have physically meanings in<br />

addition to statistical meanings in this study. Based the assumptions of the trends of mean<br />

frequency (λ) <strong>and</strong> mean intensity (µ) of wind events, we may explicitly simulate future<br />

wind conditions. Un<strong>for</strong>tunately, the scientific community is still uncertain about the fu-<br />

ture trends of tropical cyclones <strong>and</strong> other wind activities (American Meteorological Society,<br />

2007). There have been studies suggesting that tropical cyclones have become more intense<br />

over the last several decades <strong>and</strong> the trends may continue in the future (e.g. Knutson <strong>and</strong><br />

Tuleya, 2004; Elsner, 2008). The storm tracks are also projected to move, influencing the<br />

wind pattern (Intergovernmental Panel on Climate Change, 2007). However, there has not<br />

been a consensus about the degree of future changes in the intensity or frequency of wind<br />

hazards under the effect of climate change (Pielke et al., 2005). There<strong>for</strong>e, instead of look-<br />

ing at some projected climate change scenarios, we investigate a number of hypothetical<br />

scenarios in which the mean frequency (λ) <strong>and</strong> the mean intensity (µ) of wind events change<br />

linearly or exponentially over time:<br />

⎧<br />

⎪⎨ λ(t) = λ0 ± ∆λt,<br />

⎪⎩ µ(t) = µ0 ± ∆µt,<br />

⎧<br />

⎪⎨ λ(t) = λ0 exp(kλt),<br />

⎪⎩ µ(t) = µ0 ± exp(kµt),<br />

(5.9)<br />

where {λ0, ∆λ, kλ} <strong>and</strong> {µ0, ∆µ, kµ} are the initial value, yearly constant change <strong>and</strong> rate<br />

of change <strong>for</strong> the mean frequency <strong>and</strong> mean intensity, respectively.<br />

Reasonable <strong>and</strong> realistic ranges of the parameters λ <strong>and</strong> µ may be found in a study by<br />

National Institute of St<strong>and</strong>ards <strong>and</strong> Technology (NIST), which has exported the extreme<br />

wind speed data in 129 stations over the contiguous United States from another study by


Chapter 5. Long-Term <strong>Risk</strong> <strong>Assessment</strong> <strong>and</strong> <strong>Management</strong> 100<br />

Mean Intensity (!)<br />

14<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

10 1<br />

Greenville, SC<br />

Greensboro, NC<br />

10 2<br />

Portl<strong>and</strong>, OR<br />

Memphis, TN<br />

10 3<br />

Block Isl<strong>and</strong>, RI<br />

Evansville, IN<br />

Columbus, OH<br />

10 4<br />

Mean Frequency ()<br />

North Platte, NE<br />

Denver, CO<br />

10 5<br />

10 6<br />

Elkins, WV<br />

Figure 5.1: Distribution of Gumbel distribution parameters λ <strong>and</strong> µ of extreme wind speed<br />

data in 129 stations over the contiguous United States (National Institute of St<strong>and</strong>ards <strong>and</strong><br />

Technology, 2006). Some geographic locations of the stations are labeled.<br />

Simiu et al. (1979), <strong>and</strong> fitted the data into Gumbel distributions using Gumbel’s method<br />

(National Institute of St<strong>and</strong>ards <strong>and</strong> Technology, 2006a). The Gumbel distribution param-<br />

eters are converted into λ <strong>and</strong> µ <strong>and</strong> are shown in a scatter plot in Figure 5.1. Despite the<br />

wide range of values of both parameters, the 50-year-return-period wind speed only ranges<br />

from 45.4 mph to 97.2 mph fastest-mile wind. Note that the values of the parameters do not<br />

represent the actual frequency or intensity of wind hazards. For example, Figure 5.1 does<br />

not mean that Greensville, SC has much less wind events than Denver, CO. It is because the<br />

parameters are based on best-fit of the extreme wind data <strong>and</strong> do not distinguish different<br />

types of wind events. However, the underlying physical meanings of both parameters are<br />

still valid means to interpret the fitted Gumbel distribution, if we view the occurrence of<br />

wind events as a stochastic process <strong>and</strong> treat the parameters as general representations of<br />

the parameters of all types of wind events.<br />

Based on the range of parameters shown in Figure 5.1, two sets of parameters, {λ0 =<br />

10 7


Chapter 5. Long-Term <strong>Risk</strong> <strong>Assessment</strong> <strong>and</strong> <strong>Management</strong> 101<br />

Figure 5.2: Two sample results of the hazard model using (a) λ0 = 100, µ0 = 6, ∆λ =<br />

0.01λ0, ∆µ = 0.01µ0 (b) λ0 = 10, 000, µ0 = 6, ∆λ = 0.01λ0, ∆µ = 0.01µ0. Below each<br />

sample result are the wind speed distributions at the beginning <strong>and</strong> the end of the 50-year<br />

time interval.


Chapter 5. Long-Term <strong>Risk</strong> <strong>Assessment</strong> <strong>and</strong> <strong>Management</strong> 102<br />

100, µ0 = 6} <strong>and</strong> {λ0 = 10, 000, µ0 = 6}, are selected to demonstrate two sample results<br />

of the non-stationary compound Poisson process. Both sample results assume that the<br />

parameters are changing linearly with time. Comparing Figure 5.2a <strong>and</strong> 5.2b, a higher λ<br />

yields a larger extreme wind speed given the same µ.<br />

5.4 Single-Limit-State Vulnerability Model<br />

5.4.1 Model Description<br />

In structural engineering, a limit state may be defined as a set of per<strong>for</strong>mance criteria that<br />

has to be satisfied during the design process. An ultimate limit state is met if all the<br />

structural <strong>for</strong>ces are below the resistance, ensuring that the structure does not collapse in<br />

a load event. In general, a limit state may be expressed as L − R, where L is the load <strong>and</strong><br />

R is the resistance. If the limit state is smaller than zero, the structure is considered failed.<br />

Conservatively, in a hazard event, the normalized repair cost, CR, may be expressed as:<br />

⎧<br />

⎪⎨ 1, L ≥ R<br />

CR =<br />

⎪⎩ 0, L < R<br />

, (5.10)<br />

in which the structure requires full repair once the limit state is exceeded.<br />

The building code in the United States specifies that a structure has to withst<strong>and</strong> min-<br />

imally a 50-year return period wind speed. Depending on the factor of safety embedded<br />

in the structural design, a structure would normally sustain a wind speed higher than the<br />

50-year wind speed. Figure 5.3 shows three example vulnerability curves of three different<br />

factors of safety. The repair cost is a function of the factor of safety against 50-year return<br />

period wind only, <strong>and</strong> does not depend on the actual values of the 50-year wind speed or the<br />

resistance, which varies largely from region to region <strong>and</strong> from structure to structure. The<br />

simplicity of this model will allow us to illustrate the effect of climate change on structural


Chapter 5. Long-Term <strong>Risk</strong> <strong>Assessment</strong> <strong>and</strong> <strong>Management</strong> 103<br />

design without ambiguity arising from the complexity of the vulnerability model, be<strong>for</strong>e we<br />

move on to the next more complex vulnerability model.<br />

C R<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

FS = 1<br />

FS = 1.5<br />

FS = 2<br />

0 50 100 150<br />

<strong>Wind</strong> Speed (mph)<br />

Figure 5.3: Vulnerability curves of the single-limit-state deterministic model, in which the<br />

structure is designed to resist against the 50-year return period load calculated using parameters<br />

λ = 100 <strong>and</strong> µ = 6. FS denotes the factor of safety.<br />

5.4.2 Life-cycle Cost Results<br />

Consider a newly constructed structure with a 50-year design life, which is typical <strong>for</strong> general<br />

structures. The design load of the structure is calculated based on the Gumbel distribution<br />

parameters, mean intensity (µ) <strong>and</strong> mean frequency (λ) of wind events, assuming that both<br />

parameters are stationary with time throughout the design life. The factor of safety against<br />

50-year wind speed ranges from 1.5 to 3 in the following analysis, since 2 is the mean factor<br />

of safety <strong>for</strong> the ultimate limit state (Christian <strong>and</strong> Urzua, 2009), <strong>and</strong> we are interested in<br />

a higher factor of safety <strong>for</strong> adaptation to climate change. The ranges of λ <strong>and</strong> µ are [100,<br />

100,000] <strong>and</strong> [4, 10], respectively, based on the NIST Gumbel distribution parameters in<br />

Figure 5.1.<br />

Figure 5.4 shows the mean total discounted repair cost over the structure’s design life<br />

as a function of initial mean wind event frequency (λ0) <strong>and</strong> factor of safety against 50-year


(a)<br />

Change in Mean Frequency / Year<br />

(as a percentage of µ0)<br />

(b)<br />

Percentage Change in Mean Frequency / Year<br />

1.0%<br />

0.8%<br />

0.6%<br />

0.4%<br />

0.2%<br />

0%<br />

1.0%<br />

0.8%<br />

0.6%<br />

0.4%<br />

0.2%<br />

0%<br />

Chapter 5. Long-Term <strong>Risk</strong> <strong>Assessment</strong> <strong>and</strong> <strong>Management</strong> 104<br />

0%<br />

0%<br />

0.2% 0.4% 0.6% 0.8% 1.0%<br />

Change in Mean Intensity / Year<br />

(as a percentage of !0)<br />

0.2% 0.4% 0.6% 0.8% 1.0%<br />

Percentage Change in Mean Intensity / Year<br />

Initial Frequency ( 0 )<br />

10 5<br />

10 4<br />

10 3<br />

10 2<br />

LEGEND<br />

10<br />

1.5 2 2.5 3<br />

1<br />

Factor of Safety<br />

10 5<br />

10 4<br />

10 3<br />

10 2<br />

LEGEND<br />

10<br />

1.5 2 2.5 3<br />

1<br />

Factor of Safety<br />

Figure 5.4: Mean total discounted repair cost as a function of initial frequency <strong>and</strong> safety<br />

factor in 36 different climate change scenarios with (a) linearly increasing parameters (b)<br />

exponentially increasing parameters. The total discounted repair cost is denoted by the<br />

color.<br />

Initial Frequency ( 0 )<br />

0.1<br />

0.01<br />

0.001<br />

0.0001<br />

1e05<br />

1e06<br />

1e07<br />

1e08<br />

1e09<br />

0.1<br />

0.01<br />

0.001<br />

0.0001<br />

1e05<br />

1e06<br />

1e07<br />

1e08<br />

1e09


(a)<br />

Change in Mean Frequency / Year<br />

(as a percentage of µ0)<br />

(b)<br />

Percentage Change in Mean Frequency / Year<br />

1.0%<br />

0.8%<br />

0.6%<br />

0.4%<br />

0.2%<br />

0%<br />

1.0%<br />

0.8%<br />

0.6%<br />

0.4%<br />

0.2%<br />

0%<br />

Chapter 5. Long-Term <strong>Risk</strong> <strong>Assessment</strong> <strong>and</strong> <strong>Management</strong> 105<br />

0%<br />

0%<br />

0.2% 0.4% 0.6% 0.8% 1.0%<br />

Change in Mean Intensity / Year<br />

(as a percentage of !0)<br />

0.2% 0.4% 0.6% 0.8% 1.0%<br />

Percentage Change in Mean Intensity / Year<br />

Initial<br />

Initial<br />

Frequency<br />

Frequency<br />

(<br />

(<br />

)<br />

)<br />

0<br />

0<br />

Initial Frequency ( 0 )<br />

10 5 10 5<br />

10 4 10 4<br />

10 3 10 3<br />

10 2 10 2<br />

LEGEND<br />

10<br />

1.5 2 2.5 3<br />

1 10<br />

1.5 2 2.5 3<br />

Factor of Safety<br />

1<br />

Factor of Safety<br />

10 5<br />

10 4<br />

10 3<br />

10 2<br />

LEGEND<br />

10<br />

1.5 2 2.5 3<br />

1<br />

Factor of Safety<br />

1,000,000% 1,000,000<br />

10,000%<br />

100,000% 100,000%<br />

1,000%<br />

10,000%<br />

1,000%<br />

100%<br />

100%<br />

10%<br />

10%<br />

1%<br />

1%<br />

1,000,000% 1,000,000<br />

100,000%<br />

10,000%<br />

Figure 5.5: Percentage difference in mean total discounted repair costs between 36 different<br />

climate change scenarios <strong>and</strong> the ‘no climate change’ scenario. Parameters are linearly<br />

increasing with time in (a) <strong>and</strong> exponentially increasing in (b). The percentage difference<br />

is denoted by the color.<br />

1,000%<br />

100%<br />

10%<br />

1%


Chapter 5. Long-Term <strong>Risk</strong> <strong>Assessment</strong> <strong>and</strong> <strong>Management</strong> 106<br />

wind in 72 different climate change scenarios, in which λ <strong>and</strong> µ changes from 0 to 1.0% per<br />

year linearly or exponentially. Given the same value of µ, a higher λ yields a higher 50-year<br />

return-period wind speed, <strong>and</strong> hence requires a more resistant structure. Each subplot in<br />

Figure 5.4 indicates that a lower λ in fact leads to a higher expected life-cycle cost, implying<br />

that a historically safe region can have a larger expected damage than the disaster-prone<br />

regions due to the much lower resistance of structures. Also, as expected, the total repair<br />

cost is higher if the safety factor is low. These relationships are consistent throughout all<br />

the 72 scenarios.<br />

Figure 5.5 presents the percentage difference in mean total discounted repair costs be-<br />

tween the 72 different climate change scenarios <strong>and</strong> the ‘no climate change’ scenario. The<br />

life-cycle cost is much more sensitive to the change in µ than the change in λ. The mean<br />

total discounted repair costs can increase drastically if µ increases by more than 0.2% per<br />

year, despite the fact that future repair costs are being discounted by 5 % per year.<br />

In addition to the minimization of the mean life-cycle cost, a small variation of the<br />

life-cycle cost is desired in structural design optimization because it implies consistency<br />

of structural per<strong>for</strong>mance through time. Figure 5.6 shows the st<strong>and</strong>ard deviation of the<br />

total discounted repair cost in the 72 climate change scenarios. Within each scenario, the<br />

st<strong>and</strong>ard deviation decreases with λ0 <strong>and</strong> the factor of safety, which is the same as the mean<br />

in previous results. However, in all scenarios, the st<strong>and</strong>ard deviation remains within the<br />

range between 0 <strong>and</strong> 0.2, which is much smaller than that of the mean shown earlier in Figure<br />

5.4. Figure 5.7 shows the percentage difference in the coefficient of variation (ratio of the<br />

st<strong>and</strong>ard deviation to the mean) of the total repair costs between different climate change<br />

scenarios <strong>and</strong> the ‘no climate change’ scenario. In each scenario, the percentage difference<br />

in the coefficient of variation is negative, <strong>and</strong> increases (in absolute value) with λ0 <strong>and</strong> the<br />

factor of safety. Comparing different scenarios, the change in coefficient of variation is much<br />

more sensitive to the change in µ than to the change in λ. Also, the coefficient of variation


(a)<br />

Change in Mean Frequency / Year<br />

(as a percentage of µ0)<br />

(b)<br />

Percentage Change in Mean Frequency / Year<br />

1.0%<br />

0.8%<br />

0.6%<br />

0.4%<br />

0.2%<br />

0%<br />

1.0%<br />

0.8%<br />

0.6%<br />

0.4%<br />

0.2%<br />

0%<br />

Chapter 5. Long-Term <strong>Risk</strong> <strong>Assessment</strong> <strong>and</strong> <strong>Management</strong> 107<br />

0%<br />

0%<br />

0.2% 0.4% 0.6% 0.8% 1.0%<br />

Change in Mean Intensity / Year<br />

(as a percentage of !0)<br />

0.2% 0.4% 0.6% 0.8% 1.0%<br />

Percentage Change in Mean Intensity / Year<br />

Initial Frequency ( 0 )<br />

10 5<br />

10 4<br />

10 3<br />

10 2<br />

LEGEND<br />

10<br />

1.5 2 2.5 3<br />

1<br />

Factor of Safety<br />

LEGEND<br />

10<br />

1.5 2 2.5 3<br />

1<br />

Factor of Safety<br />

Figure 5.6: St<strong>and</strong>ard deviation of total discounted repair cost as a function of initial frequency<br />

<strong>and</strong> safety factor in 36 different climate change scenarios with (a) linearly increasing<br />

parameters (b) exponentially increasing parameters. The total discounted repair cost is indicated<br />

by the color.<br />

Initial Frequency ( 0 )<br />

10 5<br />

10 4<br />

10 3<br />

10 2<br />

0.2<br />

0.18<br />

0.16<br />

0.14<br />

0.12<br />

0.1<br />

0.08<br />

0.06<br />

0.04<br />

0.02<br />

0<br />

0.2<br />

0.18<br />

0.16<br />

0.14<br />

0.12<br />

0.1<br />

0.08<br />

0.06<br />

0.04<br />

0.02<br />

0


(a)<br />

Change in Mean Frequency / Year<br />

(as a percentage of µ0)<br />

(b)<br />

Percentage Change in Mean Frequency / Year<br />

1.0%<br />

0.8%<br />

0.6%<br />

0.4%<br />

0.2%<br />

0%<br />

1.0%<br />

0.8%<br />

0.6%<br />

0.4%<br />

0.2%<br />

0%<br />

Chapter 5. Long-Term <strong>Risk</strong> <strong>Assessment</strong> <strong>and</strong> <strong>Management</strong> 108<br />

0%<br />

0%<br />

0.2% 0.4% 0.6% 0.8% 1.0%<br />

Change in Mean Intensity / Year<br />

(as a percentage of !0)<br />

0.2% 0.4% 0.6% 0.8% 1.0%<br />

Percentage Change in Mean Intensity / Year<br />

Initial Frequency ( ( )<br />

0<br />

Initial Frequency ( 0 )<br />

10 5<br />

10 5<br />

10 4<br />

10 4<br />

10 3<br />

10 3<br />

10 2<br />

10 2<br />

LEGEND<br />

10<br />

1.5 2 2.5 3<br />

1<br />

10<br />

1.5 2 2.5 3<br />

Factor of Safety<br />

1<br />

Factor of Safety<br />

10 5<br />

10 4<br />

10 3<br />

10 2<br />

LEGEND<br />

10<br />

1.5 2 2.5 3<br />

1<br />

Factor of Safety<br />

10,000% 0%<br />

10%<br />

20%<br />

1,000%<br />

30%<br />

40%<br />

100% 50%<br />

60%<br />

70%<br />

10%<br />

80%<br />

90%<br />

1% 100%<br />

1,000,000%<br />

0%<br />

10%<br />

100,000%<br />

20%<br />

Figure 5.7: Percentage difference in coefficient of variation of total discounted repair costs<br />

between 36 different climate change scenarios <strong>and</strong> the ‘no climate change’ scenario. Parameters<br />

are linearly increasing with time in (a) <strong>and</strong> exponentially increasing in (b). The<br />

percentage difference is indicated by the color.<br />

30%<br />

40%<br />

50%<br />

60%<br />

70%<br />

80%<br />

90%<br />

100%


Chapter 5. Long-Term <strong>Risk</strong> <strong>Assessment</strong> <strong>and</strong> <strong>Management</strong> 109<br />

decreases as the mean total discounted repair cost increases, implying that the mean of<br />

life-cycle cost increases more rapidly than the st<strong>and</strong>ard deviation when the structure is<br />

subjected to more intense loads. This indicates that the significant increase in structural<br />

life-cycle cost, but not the uncertainty about the life-cycle cost, may be the major problem<br />

in climate change.<br />

Each climate change scenario above assumes that the changes in mean intensity (µ) or<br />

mean frequency (λ) are deterministic. However, as discussed above in the hazard model,<br />

currently the trends of future wind hazards are still uncertain. The mean <strong>and</strong> the st<strong>and</strong>ard<br />

deviation of the total discounted repair cost are recalculated, assuming that the yearly<br />

changes in λ <strong>and</strong> µ are constant <strong>and</strong> are uni<strong>for</strong>mly distributed between 0% <strong>and</strong> 0.1% of<br />

the initial values. The results are compared with the scenario in which the yearly changes<br />

are deterministically 0.05% of the initial values of parameters. The uni<strong>for</strong>mly distributed<br />

parameters yield a larger mean <strong>and</strong> a larger st<strong>and</strong>ard deviation of the total discounted<br />

repair cost, <strong>and</strong> the differences are attributed to the uncertainty in parameters. Figures<br />

5.8 <strong>and</strong> 5.9 illustrate the ratios of the total mean discounted repair cost attributed to the<br />

uncertainty in parameters with respect to that attributed to the mean value of parameters.<br />

In the most extreme case, the uncertainty in parameters can contribute additional 61% of<br />

the mean repair cost <strong>and</strong> additional 82% of the st<strong>and</strong>ard deviation of the repair cost. The<br />

contribution diminishes as the factor of safety or the initial mean wind event frequency (λ0)<br />

increases, <strong>and</strong> the degree of sensitivity to both parameters is similar.<br />

5.4.3 Life-cycle-cost-related Decisions<br />

Previous results have demonstrated different degrees of changes in total discounted repair<br />

costs in various scenarios. The followings will demonstrate how the changes in cost may<br />

impact different decisions regarding the management <strong>and</strong> design of a structure.


Initial Mean Frequency<br />

Chapter 5. Long-Term <strong>Risk</strong> <strong>Assessment</strong> <strong>and</strong> <strong>Management</strong> 110<br />

100<br />

1,000<br />

10,000<br />

100,000<br />

58%<br />

39%<br />

84%<br />

74%<br />

42%<br />

80%<br />

61% 62%<br />

Factor of Safety<br />

1.5 2.0 2.5 3.0<br />

16%<br />

26%<br />

90%<br />

5%<br />

95%<br />

10%<br />

20%<br />

38%<br />

92%<br />

79%<br />

1%<br />

99%<br />

3%<br />

97%<br />

9%<br />

21%<br />

0%<br />

100%<br />

1%<br />

99%<br />

97%<br />

89%<br />

4%<br />

11%<br />

89%<br />

89% of<br />

the mean repair cost<br />

attributed to mean<br />

value of parameters<br />

11% of the mean<br />

repair cost attributed<br />

to uncertainty in<br />

parameters<br />

Figure 5.8: Ratios of the total mean discounted repair costs attributed to the uncertainty<br />

in parameters to that attributed to the mean value of parameters.<br />

Initial Mean Frequency<br />

100<br />

1,000<br />

10,000<br />

100,000<br />

58%<br />

45%<br />

31%<br />

18%<br />

69%<br />

82%<br />

42%<br />

75%<br />

55% 63%<br />

46%<br />

28%<br />

Factor of Safety<br />

1.5 2.0 2.5 3.0<br />

25%<br />

37%<br />

54% 65%<br />

72%<br />

79%<br />

45%<br />

87%<br />

13%<br />

21%<br />

35%<br />

78%<br />

55% 60%<br />

94%<br />

88%<br />

6%<br />

12%<br />

22%<br />

40%<br />

60%<br />

11%<br />

60% of<br />

the st<strong>and</strong>ard<br />

deviation in repair cost<br />

attributed to mean value<br />

of parameters<br />

40% of the st<strong>and</strong>ard<br />

deviation in repair cost<br />

attributed to uncertainty<br />

in parameters<br />

Figure 5.9: Ratios of the st<strong>and</strong>ard deviations of the discounted repair costs attributed to<br />

the uncertainty in parameters to that attributed to the mean value of parameters.<br />

40%


Chapter 5. Long-Term <strong>Risk</strong> <strong>Assessment</strong> <strong>and</strong> <strong>Management</strong> 111<br />

5.3.3.1 Maintenance cost<br />

Structure deteriorates over time <strong>and</strong> hence requires maintenance. With resistance deterio-<br />

ration, a structure is more vulnerable to wind hazards <strong>and</strong> its life-cycle cost is expected to<br />

increase. An exponential deterioration model is applied to the study:<br />

R(t) = R0 exp(−αt), (5.11)<br />

where R0 is the initial resistance value, <strong>and</strong> α is the deterioration rate, to illustrate the<br />

maintenance-related analysis. A 0.5% deterioration rate is used, <strong>and</strong> it is assumed that<br />

after each failure, the structure is fully repaired so that the resistance is restored to the<br />

initial value in a short amount of time.<br />

Four sample simulations are shown in Figure 5.10 explaining the effect of resistance<br />

deterioration. In the simulations, the factor of safety against 50-year wind speed is 1.0.<br />

The initial mean wind event frequency (λ0) is 100 <strong>and</strong> the initial mean wind event intensity<br />

(µ0) is 6. The figure illustrates that the resistance deterioration can significantly increase<br />

the number of structural failures across the 50-year design life. If there is no change in<br />

wind event parameters over time, resistance deterioration increases the number of failures<br />

from 2 to 8. If λ <strong>and</strong> µ increase by 1% of the initial values every year over time, resistance<br />

deterioration increases the number of failures from 6 to 13. 100,000 additional simulations<br />

Table 5.1: Expected total discounted repair costs <strong>and</strong> maximum maintenance cost in two<br />

different climate change scenarios.<br />

No change in mean frequency ↑ 1% of initial frequency/yr<br />

No change in mean intensity ↑ 1% of initial intensity/yr<br />

No deterioration 0.0080 0.0333<br />

5% deterioration 0.3885 0.5746<br />

Max maintenance cost 0.3805 0.5414


No resistance<br />

deterioration<br />

5% resistance<br />

deterioration / year<br />

<strong>Wind</strong> Speed<br />

(relative to 50year speed)<br />

Resistance<br />

(relative to 50year speed)<br />

Resistance<br />

(relative to 50year speed)<br />

Chapter 5. Long-Term <strong>Risk</strong> <strong>Assessment</strong> <strong>and</strong> <strong>Management</strong> 112<br />

2<br />

1.5<br />

1<br />

0.5<br />

0<br />

0 10 20 30 40 50<br />

Time (year)<br />

2<br />

1.5<br />

1<br />

0.5<br />

0<br />

0 10 20 30 40 50<br />

Time<br />

2<br />

1.5<br />

1<br />

0.5<br />

No change in mean frequency<br />

No change in mean intensity<br />

0<br />

0 10 20 30 40 50<br />

Time<br />

<strong>Wind</strong> Speed<br />

(relative to 50year speed)<br />

Resistance<br />

(relative to 50year speed)<br />

Resistance<br />

(relative to 50year speed)<br />

"1% of initial mean frequency / year<br />

"1% of initial mean intensity / year<br />

2<br />

1.5<br />

1<br />

0.5<br />

0<br />

0 10 20 30 40 50<br />

Time (year)<br />

2<br />

1.5<br />

1<br />

0.5<br />

0<br />

0 10 20 30 40 50<br />

Time<br />

2<br />

1.5<br />

1<br />

0.5<br />

0<br />

0 10 20 30 40 50<br />

Time<br />

Figure 5.10: Four sample simulations showing the effects of resistance deterioration in two<br />

different climate scenarios. The circles indicate failure of structure.


Chapter 5. Long-Term <strong>Risk</strong> <strong>Assessment</strong> <strong>and</strong> <strong>Management</strong> 113<br />

are carried out <strong>and</strong> the resultant mean total discounted repair costs are listed in Table 5.1.<br />

The difference in life-cycle cost between the case with deterioration <strong>and</strong> the case without<br />

deterioration is the maximum maintenance cost that the owner is willing to pay over the<br />

50-year design life from the economic sense. The climate change scenario can increase the<br />

repair cost significantly <strong>and</strong> encourage maintenance of structure against deterioration.<br />

5.3.3.2 Structural Design<br />

In life-cycle-cost design, a structural design is optimized <strong>for</strong> the lowest life-cycle cost, which<br />

comprises initial construction cost, C0, <strong>and</strong> the repair cost (see equation 5.1). For simplicity,<br />

we assume that C0 is a linear function of factor safety. Figure 5.11 illustrates the total life-<br />

cycle cost as a function of factor of safety in two different climate change scenarios using<br />

λ0 = 100 <strong>and</strong> µ0 = 6. If λ <strong>and</strong> µ increase by 1% of the initial values every year over time,<br />

due to the increase in expected repair cost, a higher factor of safety is required to achieve<br />

the minimum total life-cycle cost. In this example, future climate change encourages the<br />

design of a more costly but safer structure.<br />

5.5 Modified Florida Public Hurricane Loss Projection<br />

(FPHLP) Model<br />

5.5.1 Model Description<br />

Traditionally, vulnerability models of structures are deduced from empirical relationship<br />

between wind speed <strong>and</strong> insurance claims from historical insurance loss databases. Recent<br />

developments have employed engineering theories to simulate the physical damage process<br />

when severe winds impacts a structure. Design codes <strong>and</strong> st<strong>and</strong>ards are used to calculate<br />

the wind loads during a wind hazard. Comparing the load with the resistance data of<br />

different components collected in test experiments, a building’s damage condition may be


Expected Total<br />

Discounted Repair<br />

Cost<br />

Initial Construction<br />

Cost<br />

Expected Total<br />

Life-Cycle Cost<br />

0.04<br />

0.03<br />

0.02<br />

0.01<br />

0.04<br />

0.02<br />

0.06<br />

0.04<br />

0.02<br />

Chapter 5. Long-Term <strong>Risk</strong> <strong>Assessment</strong> <strong>and</strong> <strong>Management</strong> 114<br />

No change in mean frequency<br />

No change in mean intensity<br />

0<br />

2 2.5 3 3.5 4<br />

0<br />

2 2.5 3 3.5 4<br />

Min<br />

0<br />

2 2.5 3 3.5 4<br />

0.04<br />

0.03<br />

0.02<br />

0.01<br />

0.04<br />

0.02<br />

0.06<br />

0.04<br />

0.02<br />

!1% of initial mean frequency / year<br />

!1% of initial mean intensity / year<br />

0<br />

2 2.5 3 3.5 4<br />

Factor of Safety Factor of Safety<br />

+ +<br />

Factor of Safety<br />

0<br />

2 2.5 3 3.5 4<br />

= =<br />

Factor of Safety<br />

Factor of Safety<br />

Min<br />

0<br />

2 2.5 3 3.5 4<br />

Factor of Safety<br />

Figure 5.11: Minimization of life-cycle cost in two different climate scenarios. The circle<br />

indicates the minimum life-cycle cost <strong>and</strong> the associated factor of safety.


simulated.<br />

Chapter 5. Long-Term <strong>Risk</strong> <strong>Assessment</strong> <strong>and</strong> <strong>Management</strong> 115<br />

Florida Public Hurricane Loss Projection (FPHLP) Model, which is developed by a<br />

multi-disciplinary team of experts from various universities <strong>and</strong> research institutions at<br />

Florida International University, is one of the actuarial models that were approved by the<br />

Florida Commission on Hurricane Loss Projection Methodology <strong>for</strong> estimating hurricane<br />

insurance rates in the state of Florida. Though its source code is not open, it is the most<br />

transparent model compared to the other approved models, which are proprietary <strong>and</strong><br />

commercial. Using a structural engineering approach, FPHLP model calculates the wind<br />

loads acting on a single-story house based on ASCE 7 St<strong>and</strong>ard <strong>and</strong> checks a series of limit<br />

states including the uplift <strong>for</strong>ce on roof covers <strong>and</strong> roof sheathing panels, tension of roof-to-<br />

wall connections, shear <strong>for</strong>ce on walls, <strong>and</strong> pressure on windows, doors <strong>and</strong> garage doors.<br />

The probability distribution of both the wind loads <strong>and</strong> building component resistance are<br />

explicitly specified. Using Monte Carlo simulation, the probability of various structural<br />

damage conditions under a given wind speed is calculated.<br />

We have developed a vulnerability model based on the engineering component of FPHLP<br />

model, <strong>and</strong> modified some of the mechanisms inside the model including internal pressure<br />

<strong>and</strong> wind-borne debris. The model is capable of estimating damage ratios of a range of<br />

building components (e.g. roof covers, roof panels, roof-to-wall connections, walls, windows,<br />

doors <strong>and</strong> garage doors) <strong>and</strong> the damage ratio of the structural system, which is the weighted<br />

average of the damage ratio of the building components, with the weights proportional to the<br />

repair cost of the corresponding component. Figure 5.12 shows four example vulnerability<br />

curves of different structures from the modified FPHLP model. Note that only damage of the<br />

structural system is considered in the vulnerability curves, as content damage or additional<br />

living expenses are out of the scope of an engineering model <strong>and</strong> are not considered here.


Structural Damage Ratio<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

Chapter 5. Long-Term <strong>Risk</strong> <strong>Assessment</strong> <strong>and</strong> <strong>Management</strong> 116<br />

Hip roof with tiles (no shutters)<br />

Gable roof with tiles (no shutters)<br />

Hip roof with tiles (with shutters)<br />

Hip roof with shingles (with shutters)<br />

0<br />

0 50 100 150 200 250<br />

<strong>Wind</strong> Speed (mph)<br />

Figure 5.12: Examples of vulnerability curves of the modified Florida Public Hurricane<br />

Loss Model. All the curves are related to a one-story concrete-block house with normal<br />

resistance commonly found in South Florida.<br />

5.5.2 Expected Repair Cost over the Contiguous United States<br />

Consider an one-story hip-roof concrete-block house with tiles <strong>and</strong> no shutters that is com-<br />

monly found in South Florida. Assume that the structure has a 50-year design life. Figure<br />

5.13 shows the expected total discounted repair cost of the structure in 129 geographic lo-<br />

cations over nine different combinations of percentage changes in mean intensity <strong>and</strong> mean<br />

frequency. The Gumbel distribution parameters from the extreme wind data at the 129 lo-<br />

cations are from Figure 5.1, <strong>and</strong> the wind speed data is converted from fastest mile wind to<br />

3-second gust wind, on which the modified FPHLP model is based. As observed in the last<br />

section, the results of the life-cycle costs are similar regardless of whether the parameters<br />

λ <strong>and</strong> µ are changing linearly or exponentially with time. There<strong>for</strong>e, only results of expo-<br />

nentially changing parameters are presented here due to space constraint. The expected<br />

total discounted repair costs, which are indicated by the size <strong>and</strong> the color of the markers<br />

in Figure 5.13, may be interpreted as the ratio of the repair cost due to structural damage<br />

with respect to the construction cost of the structure in the 50-year design period.<br />

In each subplot of Figure 5.13, the mean total discounted repair cost increases with<br />

initial mean frequency (λ0) <strong>and</strong> initial mean intensity (µ0). Comparing subplot to subplot,


Percentage Change in Mean Frequency / Year (k!)<br />

0.05%<br />

0%<br />

Initial Mean Intensity (! )<br />

Initial Mean Intensity (! )<br />

0 0<br />

0.1%<br />

Initial Mean Intensity (! )<br />

0<br />

14<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

14<br />

12<br />

10<br />

10 0<br />

8<br />

6<br />

4<br />

2<br />

14<br />

12<br />

10<br />

10 0<br />

8<br />

6<br />

4<br />

2<br />

10 0<br />

Chapter 5. Long-Term <strong>Risk</strong> <strong>Assessment</strong> <strong>and</strong> <strong>Management</strong> 117<br />

10 5<br />

Initial Mean Frequency ( 0 )<br />

10 5<br />

Initial Mean Frequency ( 0 )<br />

10 5<br />

Initial Mean Frequency ( 0 )<br />

Percentage Change in Mean Intensity / Year (kµ)<br />

0% 0.05% 0.1%<br />

Average:<br />

0.080<br />

Average:<br />

0.087<br />

Average:<br />

0.095<br />

10 10<br />

10 10<br />

10 10<br />

Initial Mean Intensity (! 0 )<br />

Initial Mean Intensity (! 0 )<br />

Initial Mean Intensity (! 0 )<br />

14<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

14<br />

12<br />

10<br />

10 0<br />

8<br />

6<br />

4<br />

2<br />

14<br />

12<br />

10<br />

10 0<br />

8<br />

6<br />

4<br />

2<br />

10 0<br />

10 5<br />

Initial Mean Frequency ( 0 )<br />

10 5<br />

Initial Mean Frequency ( 0 )<br />

10 5<br />

Average:<br />

0.215<br />

Average:<br />

0.244<br />

Average:<br />

0.278<br />

Initial Mean Frequency ( 0 )<br />

10 10<br />

10 10<br />

10 10<br />

Initial Mean Intensity (! 0 )<br />

1<br />

0.9<br />

0.8<br />

Initial Mean Intensity (! 0 )<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

Initial Mean Intensity (! 0 )<br />

0.3<br />

0.2<br />

0.1<br />

14<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

14<br />

12<br />

10<br />

10 0<br />

8<br />

6<br />

4<br />

2<br />

14<br />

12<br />

10<br />

10 0<br />

8<br />

6<br />

4<br />

2<br />

10 0<br />

10 5<br />

Initial Mean Frequency ( 0 )<br />

10 5<br />

Initial Mean Frequency ( 0 )<br />

10 5<br />

Average:<br />

0.792<br />

Average:<br />

0.929<br />

Average:<br />

1.091<br />

10 10<br />

10 10<br />

10 10<br />

Initial Mean Frequency () 0<br />

0<br />

0 0.2 0.4 0.6 0.8 1<br />

Figure 5.13: Expected total repair cost of a one-story hip-roof concrete-block house over 129<br />

geographic locations in contiguous United States in nine different future climate scenarios.<br />

10<br />

9<br />

8<br />

7<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0


Chapter 5. Long-Term <strong>Risk</strong> <strong>Assessment</strong> <strong>and</strong> <strong>Management</strong> 118<br />

the costs are more sensitive to the yearly percentage change in mean intensity (kµ) than<br />

the change in mean frequency (kλ), similar to the results in the last section. Also, when<br />

there is no change in mean intensity over time, the wind climate with smallest λ0 yields the<br />

largest expected repair cost. However, when kµ reaches 0.1%, the result is opposite. The<br />

wind climate with largest λ0 yields the largest expected repair cost. This indicates that a<br />

structure in a wind climate of larger µ0 is more sensitive to an increase in mean intensity<br />

of wind hazards. Overall, given a 0.1% increase in both mean intensity, the expected repair<br />

cost may increase by more than 10 times.<br />

5.5.3 Expected Repair Cost <strong>for</strong> Different <strong>Structures</strong><br />

The wind condition in Greensboro, NC (λ0 = 689.29, µ0 = 5.9743) is applied to the four<br />

vulnerability curves of different structures shown in Figure 5.12 to find out the impact of<br />

structural details on the total expected repair cost related to structural damage. The re-<br />

sults are presented in Figure 5.15 as a function of the percentage change in mean wind event<br />

intensity (kµ) only, since previous results have indicated that the repair cost’s sensitivity to<br />

kµ is significantly larger than that to the mean wind event frequency (kλ). Three compar-<br />

isons are made in the three subfigures: (a) hip roof vs. gable roof; (b) tiles vs. shingles; <strong>and</strong><br />

(c) shutters vs. no shutters. They show that the expected total repair cost may differ very<br />

significantly, from ∼20% to ∼100%, although the vulnerability looks relatively similar. A<br />

larger increase in future µ always leads to a larger difference in expected repair cost between<br />

two structures, which makes a safer structural design more financially justifiable.<br />

As mentioned above, the design codes <strong>and</strong> st<strong>and</strong>ards in the United States require build-<br />

ings to be capable to resist minimally a 50-year-return-period wind speed. However, as<br />

shown in all previous results, each parameter of the wind speed distribution does have dis-<br />

tinctive effects on a structure’s life-cycle cost in various climate scenarios. Based on the<br />

study by NIST (National Institute of St<strong>and</strong>ards <strong>and</strong> Technology, 2006a), four set of Gumbel


Expected Total Structural Damage Cost<br />

0.1<br />

0.05<br />

Hip<br />

Gable<br />

Chapter 5. Long-Term <strong>Risk</strong> <strong>Assessment</strong> <strong>and</strong> <strong>Management</strong> 119<br />

0<br />

0 0.2 0.4 0.6 0.8 1<br />

Percentage Change in Mean Intensity (k ) (%)<br />

µ<br />

Expected Total Structural Damage Cost<br />

0.1<br />

0.05<br />

No shutters<br />

Shutters<br />

0<br />

0 0.2 0.4 0.6 0.8 1<br />

Percentage Change in Mean Intensity (k ) (%)<br />

µ<br />

0<br />

0 0.2 0.4 0.6 0.8 1<br />

Percentage Change in Mean Intensity (k ) (%)<br />

µ<br />

(a) (b) (c)<br />

Expected Total Structural Damage Cost<br />

0.1<br />

0.05<br />

Tile<br />

Shingle<br />

Figure 5.14: Expected total repair cost related to structural damage: (a) hip roof vs. gable<br />

roof (b) tiles vs. shingles (c) shutters vs. no shutters.<br />

distribution parameters from four different stations (see Table 5.2) are selected <strong>and</strong> applied<br />

to two set of vulnerability curves, which represent wood-frame houses <strong>and</strong> concrete-block<br />

houses at different ages.<br />

The left of Figure 5.15 shows the vulnerability curves of three concrete-block houses<br />

that are commonly found in North Florida <strong>and</strong> are built be<strong>for</strong>e 1970, between 1970 - 1993,<br />

<strong>and</strong> after 1994. They are labeled as ‘weak’, ‘medium’, <strong>and</strong> ‘strong’, respectively. The right<br />

of the figure shows the expected total repair cost as a function of the percentage change<br />

in mean wind event intensity (kµ) in the four geographic locations. The mean frequency λ<br />

is assumed to have no change over time. Although the 50-year wind speed is almost the<br />

Table 5.2: Gumbel distribution parameters of four different geographic locations with similar<br />

50-year wind speed.<br />

Location Mean frequency (λ0) Mean intensity (µ0) 50-year wind speed (mph)<br />

Evansviille, IN 11108.1 4.699 62.11<br />

Denver, CO 244592.5 3.815 62.23<br />

Greensboro, NC 689.3 5.974 62.36<br />

Memphis, TN 4169.6 5.103 62.44


Chapter 5. Long-Term <strong>Risk</strong> <strong>Assessment</strong> <strong>and</strong> <strong>Management</strong> 120<br />

same in the four locations, the structure can have four significantly different expected total<br />

discounted repair costs over its design life. And such difference increases as kµ increases.<br />

The same applies to wood-frame houses in Figure 5.16, except that the expected total<br />

discounted repair costs are less sensitive to the age of the structure.<br />

Structural Damage Ratio<br />

1<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

Vulnerability Curves<br />

Weak<br />

Medium<br />

Strong<br />

0<br />

0 50 100 150 200 250<br />

<strong>Wind</strong> Speed (mph)<br />

Mean Total<br />

Mean Structural Total Structural Damage Damage Cost Cos<br />

Mean Total<br />

Mean Structural Total Structural Damage Damage Cost Cos<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

0 0.2 0.4 0.6 0.8 1<br />

Percentage Change in Mean Intensity (k ) (%)<br />

µ<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

Evansville, IN Denver, CO<br />

0<br />

0 0.2 0.4 0.6 0.8 1<br />

Percentage Change in Mean Intensity (k ) (%)<br />

µ<br />

Mean Total<br />

Mean Structural Total Structural Damage Damage Cost Cos<br />

Mean Total<br />

Mean Structural Total Structural Damage Damage Cost Cos<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

0 0.2 0.4 0.6 0.8 1<br />

Percentage Change in Mean Intensity (k ) (%)<br />

µ<br />

Greensboro, NC Memphis TN<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

0 0.2 0.4 0.6 0.8 1<br />

Percentage Change in Mean Intensity (k ) (%)<br />

µ<br />

Figure 5.15: Vulnerability curves <strong>and</strong> total expected repair cost related to structural damage<br />

of a concrete-block house at four different geographic locations.<br />

5.6 Closing Remarks<br />

This study has examined some of the possible effects of climate change on the life-cycle<br />

cost of structures from a wind engineering aspect. A hazard model was developed by<br />

converting the Gumbel distribution into a compound Poisson process to simulate future<br />

wind conditions. The hypothetical climate assumptions are made to calculate the life-cycle<br />

cost of structures using two vulnerability models: a single-limit-state vulnerability model<br />

<strong>and</strong> the modified FPHLP model.<br />

The results in this study have illustrated that future trends in wind hazards can affect


Structural Damage Ratio<br />

1<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

Weak<br />

Medium<br />

Strong<br />

Chapter 5. Long-Term <strong>Risk</strong> <strong>Assessment</strong> <strong>and</strong> <strong>Management</strong> 121<br />

Vulnerability Curves<br />

0<br />

0 50 100 150 200 250<br />

<strong>Wind</strong> Speed (mph)<br />

Mean Total<br />

Mean Structural Total Structural Damage Damage Cost Cos<br />

Mean Total<br />

Mean Structural Total Structural Damage Damage Cost Cos<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

0 0.2 0.4 0.6 0.8 1<br />

Percentage Change in Mean Intensity (k ) (%)<br />

µ<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

Evansville, IN Denver, CO<br />

0<br />

0 0.2 0.4 0.6 0.8 1<br />

Percentage Change in Mean Intensity (k ) (%)<br />

µ<br />

Mean Total<br />

Mean Structural Total Structural Damage Damage Cost Cos<br />

Mean Total<br />

Mean Structural Total Structural Damage Damage Cost Cos<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

0 0.2 0.4 0.6 0.8 1<br />

Percentage Change in Mean Intensity (k ) (%)<br />

µ<br />

Greensboro, NC Memphis TN<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

0 0.2 0.4 0.6 0.8 1<br />

Percentage Change in Mean Intensity (k ) (%)<br />

µ<br />

Figure 5.16: Vulnerability curves <strong>and</strong> total expected repair cost related to structural damage<br />

of a wood-frame house at four different geographic locations.<br />

the expected future repair cost of structures very significantly. Given the uncertainty of<br />

future climate trends, the future risk of structural failures remain ambiguous. This can<br />

remarkably affect the insurance market, since insurers would charge 25% higher premiums<br />

<strong>for</strong> ambiguous risks than <strong>for</strong> risks with probabilities that were well specified (Kunreuther<br />

<strong>and</strong> Michel-Kerjan, 2009).<br />

As illustrated in the results of the modified FPHLP model, different types of structural<br />

design can have very different life-cycle cost in long-term, particularly if the future wind<br />

hazards intensify, although they may have similar vulnerability curves. Currently, insurance<br />

companies have only limited economic incentive to set individual insurance rates based on<br />

the risk characteristics of each property. More in<strong>for</strong>mation on each property can help reduce<br />

the ambiguity of risk <strong>and</strong> hence the insurance premium.<br />

In addition, the life-cycle cost depends significantly on the wind speed distribution pa-<br />

rameters rather than the design wind speed. If the effects of climate change are considered,


Chapter 5. Long-Term <strong>Risk</strong> <strong>Assessment</strong> <strong>and</strong> <strong>Management</strong> 122<br />

a life-cycle cost design should investigate the actual wind speed distribution as well as its<br />

possible changes in the future. Traditional design only considers the 50-year wind speed<br />

based on the wind map in design codes <strong>and</strong> st<strong>and</strong>ards <strong>and</strong> assumes that the environmental<br />

loads are stationarity with time in long term.


Chapter<br />

6<br />

Conclusions <strong>and</strong> Future Work<br />

6.1 Summary <strong>and</strong> Conclusions<br />

This research first investigates the wind vulnerability of structural components. The method-<br />

ology is then extended to the analysis of entire residential structures <strong>and</strong> hence communities<br />

of residential structures. The last part of the dissertation extends the analysis to the long-<br />

term risk assessment <strong>and</strong> management of structures under the effect of climate change. The<br />

following highlights the findings <strong>and</strong> concludes each of the chapters:<br />

Chapter 2 compares two prevalent methodologies to specify loads <strong>and</strong> estimate damage<br />

<strong>for</strong> low-rise structures due to tropical cyclones. One method, the time-stepping method,<br />

evaluates structural loads <strong>and</strong> per<strong>for</strong>mance at consecutive time steps, accounting <strong>for</strong> evolv-<br />

ing wind direction during an individual wind event. The other, commonly known as point-<br />

in-time reliability analysis, approximates structural per<strong>for</strong>mance in an event by referencing<br />

to the events maximum wind speed only. To compare the two approaches, a database of<br />

synthetic tropical-cyclone wind velocity time series is generated based on a probabilistic<br />

model. A test set of reliability analyses indicates that the point-in-time method, by ignor-<br />

ing in<strong>for</strong>mation about directionality, underestimates the maximum wind load on structures.<br />

123


Chapter 6. Conclusions <strong>and</strong> Future Work 124<br />

The differences tend to grow (1) the greater the variability of the pressure coefficient with<br />

respect to wind direction; or (2) when the structure is within the tropical cyclones radius of<br />

maximum wind <strong>for</strong> some time period; or (3) when events intensity is relatively low. The re-<br />

sults also show that the probability of failure approximated using the point-in-time method<br />

may be underestimated by up to 60%.<br />

Chapter 3 compares three different internal pressure adjustment methods (FPHLP<br />

model, ASCE 7 St<strong>and</strong>ard, <strong>and</strong> Eurocode), <strong>and</strong> examines each of their significance in the<br />

structural vulnerability analysis. During major wind events, the presence of an opening<br />

on a building envelope may result in high internal pressure <strong>and</strong> hence a critical increase in<br />

the wind load on the roof. Due to the complexity of aerodynamics, there has not been a<br />

consensus on how to determine the value of internal pressure numerically given the damage<br />

of the building envelope. This study illustrates the interplay between internal pressure <strong>and</strong><br />

building envelope damage through several numerical examples, <strong>and</strong> presents the roof dam-<br />

age results of three baseline structures using different internal pressure adjustment methods<br />

<strong>and</strong> wind-borne debris parameters over a wide range of wind speeds. It is found that the<br />

internal pressure significantly alters the roof damage, but the degree of effect depends on<br />

the details of the internal pressure adjustment model. Additional windows, presence of<br />

wind-borne debris, <strong>and</strong> increase in wind speed does not necessarily exacerbate the roof<br />

damage in every internal pressure model. Depending on the wind speed <strong>and</strong> the structural<br />

configuration, the damage results may differ by a few percent to a few hundred percent.<br />

Chapter 4 presents an integrated vulnerability model to estimate structural damage <strong>and</strong><br />

related economic losses in clusters of residential buildings due to tropical cyclone winds. The<br />

model couples a component-based pressure damage model, which is based on the FPHLP<br />

model, <strong>and</strong> a wind-borne debris model by Lin <strong>and</strong> Vanmarcke (2008, 2010), accounting <strong>for</strong><br />

the occurrence of a possible ‘chain reaction’ of events involving wind pressure damage <strong>and</strong><br />

wind-borne debris damage, amplifying overall losses. The first part of the chapter presents


Chapter 6. Conclusions <strong>and</strong> Future Work 125<br />

the details of the methodology. The second part (1) illustrates how a structure is affected by<br />

neighboring buildings resistance against high wind; (2) contrasts the difference in damage<br />

predictions between time series analysis <strong>and</strong> point-in-time analysis during a storm passage;<br />

<strong>and</strong> (3) extends the estimation of wind-induced losses from structural components to the<br />

entire structural system, the interior system, <strong>and</strong> the utility system (including electrical,<br />

pluming, heating systems) of buildings. Different approximate damage <strong>and</strong> loss prediction<br />

methods are compared, <strong>and</strong> four numerical examples are provided.<br />

Chapter 5 studies the potential impact of climate change on structural engineering<br />

analysis, particularly the life-cycle cost of structures. A non-stationary stochastic process<br />

is developed based on the assumption that extreme wind speed is Gumbel distributed. The<br />

process simulates the long-term wind climate assuming that the mean frequency <strong>and</strong>/or<br />

the mean intensity of future wind events change linearly or exponentially with time. The<br />

wind event simulation results are applied to two vulnerability models: a single-limit-state<br />

vulnerability model <strong>and</strong> the modified FPHLP model. The life-cycle cost of a structure<br />

over its design life is calculated over a range of wind climate <strong>and</strong> a number of hypothetical<br />

climate change scenarios. The results show that future trends in wind hazards may affect<br />

the expected repair cost of a structure very significantly: a 0.1% or 0.2% yearly increase<br />

in mean intensity of wind hazards may double or triple the expected repair cost. Also,<br />

uncertainty in future wind intensity <strong>and</strong> frequency can almost double the mean <strong>and</strong> the<br />

st<strong>and</strong>ard deviation of the future repair cost.<br />

6.2 Future Work<br />

6.2.1 Structural Components<br />

The results in Chapter 2 show that the damage level of structural components is signifi-<br />

cantly affected by the trend <strong>and</strong> variability of the tropical cyclone wind time history. The


Chapter 6. Conclusions <strong>and</strong> Future Work 126<br />

effect of such trend <strong>and</strong> variability may be quantified, so that general guidelines can be<br />

developed regarding how much damage should be expected given certain characteristics of<br />

the tropical cyclone wind (besides maximum wind speed). First, one or several measures<br />

may be developed to quantify the shape or trend of the wind time history. Examples are<br />

the total change in wind angle, the weighted change in wind angle (weights proportional to<br />

wind speed), etc. Relationships between such measures <strong>and</strong> the structural damage may be<br />

obtained through numerical simulations. The goal is to express wind damage as a function<br />

of not only the maximum wind speed over time, but also other measures which describes<br />

the wind time history.<br />

Another potential future work is regarding the uncertainty of the pressure coefficient,<br />

which is a critical component of the wind load equation (see Equation 2.12). Chapter 2<br />

obtains the pressure coefficient functions from the wind tunnel experiments <strong>and</strong> other vul-<br />

nerability models, <strong>and</strong> does not take into account the uncertainty of the pressure coefficient<br />

itself. Depending on the wind speed, wind directionality <strong>and</strong> the building component, the<br />

degree of uncertainty of the pressure coefficient may change. An uncertainty analysis may<br />

be carried out to examine the effect of changing wind directionality over time on the uncer-<br />

tainty of the damage level of building components. Any change in uncertainty may affect<br />

the probability of failure of the building components.<br />

6.2.2 Low-Rise <strong>Structures</strong><br />

Chapter 3 has shown that the details of the internal pressure adjustment mechanisms may<br />

completely change the results of the roof damage estimation. Such analysis may be extended<br />

to other parts of the structure <strong>and</strong> hence the economic loss estimation, further illustrating<br />

the importance of the internal pressure adjustment mechanism. Similarly, different internal<br />

pressure adjustment mechanisms may be applied to the vulnerability analysis of multi-<br />

structural community. This may further highlight the importance of the interplay between


Chapter 6. Conclusions <strong>and</strong> Future Work 127<br />

change in internal pressure <strong>and</strong> wind-borne debris risk. Furthermore, by comparing the<br />

damage surveys with the results of different internal pressure adjustment methods, one<br />

may conclude which internal pressure adjustment method better reflects the reality.<br />

More importantly, the study presented in Chapter 3 indicates the need of wind tunnel<br />

experiments regarding internal pressure changes inside a damaged structure. While the<br />

internal pressure determination method differs very much in different design specifications<br />

<strong>and</strong> loss models, there is no consensus in the field regarding the how to determine the inter-<br />

nal pressure value numerically based on the damage on building envelope. Also, there are no<br />

publicly available results of wind tunnel experiments regarding this issue. Further investi-<br />

gations using wind tunnel <strong>and</strong> <strong>and</strong> full-scale wind experiments are necessary be<strong>for</strong>e deriving<br />

an accurate numerical method to determine internal pressure changes due to openings on<br />

the building envelope.<br />

6.2.3 Multi-Structural Communities<br />

Chapter 4 presents the integrated vulnerability model which couples the debris risk model<br />

<strong>and</strong> the pressure damage model, <strong>and</strong> applies the model to communities of residential struc-<br />

tures. In the wind vulnerability analysis of a multi-structural community, wind load may<br />

be significantly affected by the shielding effect due to the presence of nearby buildings <strong>and</strong><br />

trees. A more complete analysis should consider the shielding effect as well as possible effect<br />

of topography on the wind load acting on each structure. Such analysis may reveal a better<br />

layout of the residences in order to minimize the wind loads, which is similar to finding out<br />

the best layout of wind turbines in order to maximize the energy output of a wind farm.<br />

The third numerical example in Chapter 4 has shown the effects of a house’s strength<br />

on the other house’s wind vulnerability. The analysis may be extended to further analyze<br />

the effect of a weak house on the entire neighborhood. A more detailed analysis which<br />

reveals the economic cost of the presence of a deteriorated house will be very useful to the


Chapter 6. Conclusions <strong>and</strong> Future Work 128<br />

insurance industry <strong>and</strong> the homeowners in identifying the potential sources of danger in a<br />

residential community.<br />

6.2.4 Low-Term <strong>Risk</strong> <strong>Assessment</strong><br />

The analysis in Chapter 5 may be enhanced <strong>and</strong> extended in a number of areas. For example,<br />

Chapter 5 develops a compound Poisson process to simulate future wind events. Instead,<br />

more sophisticated <strong>and</strong> complex stochastic processes may be developed. Different wind<br />

hazards may be considered individually <strong>and</strong> the stochastic processes may reflect periodic<br />

environments like those <strong>for</strong>ming tropical cyclones (Cook et al., 2003; Lu <strong>and</strong> Garrido, 2005).<br />

Furthermore, in Chapter 5, the hypothetical climate change scenarios are limited to linear<br />

<strong>and</strong> exponential increase in wind hazard intensity <strong>and</strong> frequency of occurrence. A more<br />

complex assumption may be made, i.e. only certain levels of wind hazard events increase in<br />

frequency in the future.<br />

Also, a more detailed analysis on the life-cycle cost may be conducted. For instance,<br />

the tails of the probability distribution of repair costs, in addition to the mean <strong>and</strong> the<br />

st<strong>and</strong>ard deviation, may be analyzed in details, since they are essential in catastrophe<br />

management. An uncertainty analysis that attributes the sources of uncertainty in life-<br />

cycle cost to different origins, including hazard model, vulnerability model <strong>and</strong> life-cycle<br />

cost <strong>for</strong>mulations, will be very insightful in guiding the future developments of the risk<br />

assessment related to wind hazards.


Bibliography<br />

American Meteorological Society (2007). Climate change. Bulletin of the American Mete-<br />

orological Society, 88(418).<br />

American Society of Civil Engineers (2003). Minimum design loads <strong>for</strong> buildings <strong>and</strong> other<br />

structures, SEI/ASCE 7-02. American Society of Civil Engineers, Reston, VA.<br />

Blake, E. S., Rappaport, E. N., <strong>and</strong> L<strong>and</strong>sea, C. W. (2007). The deadliest, costliest, <strong>and</strong><br />

most intense United States tropical cyclones from 1851 to 2006 (<strong>and</strong> other frequently<br />

requested hurricane facts). Technical Report NWS TPC-5, National Weather Service,<br />

National Hurricane Center, Miami, Florida.<br />

Ch<strong>and</strong>ler, A., Jones, E., <strong>and</strong> Patel, M. (2001). Property loss estimation <strong>for</strong> wind <strong>and</strong><br />

earthquake perils. <strong>Risk</strong> Analysis, 21(2):235–249.<br />

Chang, S. <strong>and</strong> Shinozuka, M. (1996). Life-cycle cost analysis with natural hazard risk.<br />

Journal of Infrastructure Systems, 2(3):118–125.<br />

Chiu, G. L. F. (1994). An extreme-wind risk assessment system. PhD thesis, Department<br />

of Civil Engineering, Stan<strong>for</strong>d University, Stan<strong>for</strong>d, Cali<strong>for</strong>nia.<br />

Christian, J. <strong>and</strong> Urzua, A. (2009). Reliability related to factor of safety <strong>and</strong> uncertainty.<br />

In Isk<strong>and</strong>er, M., Laefer, D. F., <strong>and</strong> Hussein, M. M., editors, Contemporary Topics in In<br />

Situ Testing, Analysis, <strong>and</strong> Reliability of Foundations (GSP 186), pages 372–378, Reston,<br />

VA. ASCE.<br />

129


BIBLIOGRAPHY 130<br />

Civil <strong>and</strong> Coastal Engineering Department of the University of Florida (2009). Florida<br />

coastal monitoring program. http://fcmp.ce.ufl.edu/.<br />

Coffelt, D., Hendrickson, C., <strong>and</strong> Healey, S. (2010). Inspection, condition assessment, <strong>and</strong><br />

management decisions <strong>for</strong> commercial roof systems. Journal of Architectural Engineering,<br />

16(3):94–99.<br />

Cook, N., Harris, R., <strong>and</strong> Whiting, R. (2003). Extreme wind speeds in mixed climates<br />

revisited. Journal of <strong>Wind</strong> Engineering & Industrial Aerodynamics, 91(3):403–422.<br />

Davidson, R. A. <strong>and</strong> Rivera, M. C. (2003). Projecting building inventory changes <strong>and</strong> the<br />

effect on hurricane risk. Journal of Urban Planning <strong>and</strong> Development, 129(4):211–230.<br />

Ellingwood, B., Rosowsky, D., Li, Y., <strong>and</strong> Kim, J. (2004). Fragility assessment of light-<br />

frame wood construction subjected to wind <strong>and</strong> earthquake hazards. Journal of Structural<br />

Engineering, 130(12):1921–1930.<br />

Ellingwood, B. <strong>and</strong> Tekie, P. (1999). <strong>Wind</strong> load statistics <strong>for</strong> probability-based structural<br />

design. Journal of Structural Engineering, 125(4):453–463.<br />

Elsner, J. B. (2008). Hurricanes <strong>and</strong> climate change. Bulletin of the American Meteorological<br />

Society, 89(5):677–679.<br />

EM-DAT (2010). The international disaster database, Centre <strong>for</strong> Research on the Epidemi-<br />

ology of Disasters. http://www.emdat.be/database.<br />

Embrechts, P., Klüppelberg, C., <strong>and</strong> Mikosch, T. (1997). Modelling extremal events <strong>for</strong><br />

insurance <strong>and</strong> finance. Springer.<br />

European Committee <strong>for</strong> St<strong>and</strong>ardization/Technical committee (CEN/TC 250) (2004).<br />

General actions - Part 1-4: <strong>Wind</strong> actions. In Eurocode 1: Actions on structures. prEN<br />

1991-1-4, CEN TC 250.


BIBLIOGRAPHY 131<br />

Federal Emergency <strong>Management</strong> Agency (2006). HAZUS-MH MR3 hurricane model tech-<br />

nical manual. Federal Emergency <strong>Management</strong> Agency.<br />

Frangopol, D. M. <strong>and</strong> Maute, K. (2003). Life-cycle reliability-based optimization of civil<br />

<strong>and</strong> aerospace structures. Computers & <strong>Structures</strong>, 81(7):397–410.<br />

Fritz, W., Bienkiewicz, B., Cui, B., Flam<strong>and</strong>, O., Ho, T., Kikitsu, H., Letch<strong>for</strong>d, C.,<br />

<strong>and</strong> Simiu, E. (2008). International comparison of wind tunnel estimates of wind ef-<br />

fects on low-rise buildings: Test-related uncertainties. Journal of Structural Engineering,<br />

134(12):1887–1890.<br />

Galambos, J. <strong>and</strong> Macri, N. (1999). Classical extreme value model <strong>and</strong> prediction of extreme<br />

winds. Journal of Structural Engineering, 125(7):792–794.<br />

Ginger, J. (2000). Internal pressures <strong>and</strong> cladding net wind loads on full-scale low-rise<br />

building. Journal of Structural Engineering, 126(4):538–543.<br />

Ginger, J., Mehta, K., <strong>and</strong> Yeatts, B. (1997). Internal pressures in a low-rise full-scale<br />

building. Journal of <strong>Wind</strong> Engineering <strong>and</strong> Industrial Aerodynamics, 72(1-3):163–174.<br />

Gurley, K., Pinelli, J.-P., Subramanian, C., Cope, A., Zhang, L., Murphree, J., Artiles, A.,<br />

Misra, P., Gulati, S., <strong>and</strong> Simiu, E. (2005a). Florida Public Hurricane Loss Projection<br />

Model engineering team final report volume I: Exposure <strong>and</strong> vulnerability components<br />

of the Florida Public Hurricane Loss Projection Model. Technical report, International<br />

Hurricane Research Center, Florida International University.<br />

Gurley, K., Pinelli, J.-P., Subramanian, C., Cope, A., Zhang, L., Murphree, J., Artiles, A.,<br />

Misra, P., Gulati, S., <strong>and</strong> Simiu, E. (2005b). Florida Public Hurricane Loss Projection<br />

Model engineering team final report volume II: Predicting the vulnerability of typical<br />

residential buildings to hurricane damage. Technical report, International Hurricane Re-<br />

search Center, Florida International University.


BIBLIOGRAPHY 132<br />

Gurley, K., Pinelli, J.-P., Subramanian, C., Cope, A. D., Zhang, L., Murphree, J., Arnoldo,<br />

A., <strong>and</strong> Pranay, M. (2006). Florida Public Hurricane Loss Projection Model engineer-<br />

ing team final report volume III: Development calibration <strong>and</strong> validation of vulnerability<br />

matrices of the Florida Public Hurricane Loss Projection Model. Technical report, Inter-<br />

national Hurricane Research Center, Florida International University.<br />

Ho, T., Surry, D., Morrish, D., <strong>and</strong> Kopp, G. A. (2005). The UWO contribution to the<br />

NIST aerodynamic database <strong>for</strong> wind loads on low buildings: Part 1. Archiving <strong>for</strong>mat<br />

<strong>and</strong> basic aerodynamic data. Journal of <strong>Wind</strong> Engineering <strong>and</strong> Industrial Aerodynamics,<br />

93(1):1–30.<br />

Holl<strong>and</strong>, G. J. (1980). An analytic model of the wind <strong>and</strong> pressure profiles in hurricanes.<br />

Monthly Weather Review, 108(8):1212–1218.<br />

Holmes, J. (1996). Vulnerability curves <strong>for</strong> buildings in tropical-cyclone regions. In Fran-<br />

gopol, D. M. <strong>and</strong> Grigoriu, M. D., editors, Probabilistic Mechanics <strong>and</strong> Structural Relia-<br />

bility, pages 78–81.<br />

Holmes, J. D. (2007). <strong>Wind</strong> loading of structures. Taylor <strong>and</strong> Francis, second edition.<br />

Howard, R. A., Matheson, J., <strong>and</strong> North, W. (1972). The decision to seed hurricanes.<br />

Science, 176(4040):1191–1202.<br />

Huang, Z., Rosowsky, D. V., <strong>and</strong> Sparks, P. R. (2001). Hurricane simulation techniques <strong>for</strong><br />

the evaluation of wind-speeds <strong>and</strong> expected insurance losses. Journal of <strong>Wind</strong> Engineering<br />

<strong>and</strong> Industrial Aerodynamics, 89(7-8):605–617.<br />

Hubbert, G. D., Holl<strong>and</strong>, G. J., Leslie, L. M., <strong>and</strong> Manton, M. J. (1991). A real-time system<br />

<strong>for</strong> <strong>for</strong>ecasting tropical cyclone storm surges. Weather <strong>and</strong> Forecasting, 6(1):86–97.<br />

Hundecha, Y., St-Hilaire, A., Ouarda, T., Adlouni, S. E., <strong>and</strong> Gachon, P. (2008). A nonsta-<br />

tionary extreme value analysis <strong>for</strong> the assessment of changes in extreme annual wind speed


BIBLIOGRAPHY 133<br />

over the gulf of St. Lawrence Canada. Journal of Applied Meteorology <strong>and</strong> Climatology,<br />

47(11):2745–2759.<br />

Iman, R. L., Johnson, M. E., <strong>and</strong> Schroeder, T. A. (2002). Assessing hurricane effects. part<br />

1. sensitivity analysis. Reliability Engineering & System Safety, 78(2):131–145.<br />

Intergovernmental Panel on Climate Change (2007). Fourth <strong>Assessment</strong> Report: Climate<br />

Change 2007: The AR4 Synthesis Report. Geneva: IPCC.<br />

Jakobsen, F. <strong>and</strong> Madsen, H. (2004). Comparison <strong>and</strong> further development of parametric<br />

tropical cyclone models <strong>for</strong> storm surge modelling. Journal of <strong>Wind</strong> Engineering <strong>and</strong><br />

Industrial Aerodynamics, 92(5):375–391.<br />

Kasperski, M. (1998). Climate change <strong>and</strong> design wind load concepts. <strong>Wind</strong> <strong>and</strong> <strong>Structures</strong>,<br />

1(2):145–160.<br />

Katz, R. (2002). Stochastic modeling of hurricane damage. Journal of Applied Meteorology,<br />

41(7):754–762.<br />

Kendall, M., Stuart, A., Ord, K. J., <strong>and</strong> Arnold, S. (1999). Kendall’s Advanced Theory of<br />

Statistics: Volume 2A. A Hodder Arnold Publication, sixth edition.<br />

Kh<strong>and</strong>uri, A. C. <strong>and</strong> Morrow, G. C. (2003). Vulnerability of Buildings to <strong>Wind</strong>storms <strong>and</strong><br />

Insurance Loss Estimation. Journal of <strong>Wind</strong> Engineering <strong>and</strong> Industrial Aerodynamics,<br />

91(4):455–467.<br />

Knabb, R. D., Rhome, J. R., <strong>and</strong> Brown, D. P. (2005). Hurricane Katrina: Tropical Cyclone<br />

Report. Technical report, National Hurricane Center.<br />

Knutson, T. <strong>and</strong> Tuleya, R. (2004). Impact of co2-induced warming on simulated hurricane<br />

intensity <strong>and</strong> precipitation: Sensitivity to the choice of climate model <strong>and</strong> convective<br />

parameterization. Journal of Climate, 17(18):3477–3495.


BIBLIOGRAPHY 134<br />

Kopp, G. A., Oh, J. H., <strong>and</strong> Inculet, D. R. (2008). <strong>Wind</strong>-induced internal pressures in<br />

houses. Journal of Structural Engineering, 134(7):1129–1138.<br />

Kunreuther, H. <strong>and</strong> Michel-Kerjan, E. (2009). At War with the Weather: Managing Large-<br />

Scale <strong>Risk</strong>s in a New Era of Catastrophes. The MIT Press.<br />

Lee, K. H. <strong>and</strong> Rosowsky, D. (2005). Fragility assessment <strong>for</strong> roof sheathing failure in high<br />

wind regions. Engineering <strong>Structures</strong>, 27(6):857–868.<br />

Leicester, R. (1981). A risk model <strong>for</strong> cyclone damage to dwellings. In 3rd International<br />

Conference on Structural Safety <strong>and</strong> Reliability, pages 761–771, Trondheim, Norway.<br />

Leicester, R., Bubb, C., Dorman, C., <strong>and</strong> Beres<strong>for</strong>d, F. (1979). An assessment of potential<br />

cyclone damage to dwellings in australia. In Cermak, J. E., editor, Proceedings of the<br />

Fifth International Conference on <strong>Wind</strong> Engineering, pages 23–26, Fort Collins, New<br />

York. Pergamon Press.<br />

Li, Y. <strong>and</strong> Ellingwood, B. (2006). Hurricane damage to residential construction in the<br />

US: Importance of uncertainty modeling in risk assessment. Engineering <strong>Structures</strong>,<br />

28(7):1009–1018.<br />

Lin, N., Letch<strong>for</strong>d, C., <strong>and</strong> Holmes, J. (2006). Investigation of plate-type windborne de-<br />

bris. part i. experiments in wind tunnel <strong>and</strong> full scale. Journal of <strong>Wind</strong> Engineering &<br />

Industrial Aerodynamics, 94(2):51–76.<br />

Lin, N. <strong>and</strong> Vanmarcke, E. (2008). <strong>Wind</strong>borne debris risk assessment. Probabilist Engi-<br />

neering Mechanics, 23(4):523–530.<br />

Lin, N. <strong>and</strong> Vanmarcke, E. (2010). <strong>Wind</strong>borne debris risk analysis - Part I. Introduction<br />

<strong>and</strong> methodology. <strong>Wind</strong> <strong>and</strong> <strong>Structures</strong>, 13(2):191–206.


BIBLIOGRAPHY 135<br />

Lin, N., Vanmarcke, E., <strong>and</strong> Yau, S. C. (2010). <strong>Wind</strong>borne debris risk analysis - Part II.<br />

Application to structural vulnerability modeling. <strong>Wind</strong> <strong>and</strong> <strong>Structures</strong>, 13(2):207–220.<br />

Lu, Y. <strong>and</strong> Garrido, J. (2005). Doubly periodic non-homogeneous poisson models <strong>for</strong> hur-<br />

ricane data. Statistical Methodology, 2(1):17–35.<br />

Malmquist, D. L. <strong>and</strong> Michaels, A. F. (2000). Severe storms <strong>and</strong> the insurance industry.<br />

In R. A. Pielke, J. <strong>and</strong> R. A. Pielke, S., editors, Storms, chapter 5, pages 54–69. Sr.<br />

Routledge Press.<br />

McDonald, J. (1990). Impact resistance of common building materials to tornado missiles.<br />

Journal of <strong>Wind</strong> Engineering <strong>and</strong> Industrial Aerodynamics, 36(1-3):717–723.<br />

Minor, J., Harris, P., <strong>and</strong> Beason, W. (1978). Designing <strong>for</strong> windborne missiles in urban<br />

areas. Journal of the Structural Division, 104(11):1749–1760.<br />

Minor, J. <strong>and</strong> Mehta, K. (1979). <strong>Wind</strong> Damage Observations <strong>and</strong> Implications. Journal of<br />

the Structural Division, 105:2279–2291.<br />

Mitsuta, Y., Fujii, T., <strong>and</strong> Nagashima, I. (1996). A predicting method of typhoon wind<br />

damages. In Frangopol, D. M., editor, Probabilistic Mechanics <strong>and</strong> Structural Reliability,<br />

pages 970–973, New York. ASCE.<br />

Naess, A. <strong>and</strong> Gaidai, O. (2009). Estimation of extreme values from sampled time series.<br />

Structural Safety, 31(4):325–334.<br />

National Association of Home Builders (1999). An industry perspective on per<strong>for</strong>mance<br />

guidelines <strong>for</strong> structural safety <strong>and</strong> serviceability of one <strong>and</strong> two-family dwellings. Tech-<br />

nical report, Research Center Rep., National Association of Home Builders, Upper Marl-<br />

boro, MD.


BIBLIOGRAPHY 136<br />

National Institute of St<strong>and</strong>ards <strong>and</strong> Technology (2006a). Extreme wind speeds software:<br />

Excel. http://http://www.itl.nist.gov/div898/winds/excel.htm.<br />

National Institute of St<strong>and</strong>ards <strong>and</strong> Technology (2006b). Per<strong>for</strong>mance of physical struc-<br />

tures in hurricane katrina <strong>and</strong> hurricane rita: A reconnaissance report. Technical report,<br />

National Institute of St<strong>and</strong>ards <strong>and</strong> Technology.<br />

Oh, J. H., Kopp, G. A., <strong>and</strong> Inculet, D. R. (2007). The UWO contribution to the NIST<br />

aerodynamic database <strong>for</strong> wind loads on low buildings: Part 3. Internal pressures. Journal<br />

of <strong>Wind</strong> Engineering <strong>and</strong> Industrial Aerodynamics, 95(8):755–779.<br />

Pielke, R., L<strong>and</strong>sea, C., Mayfield, M., Laver, J., <strong>and</strong> Pasch, R. J. (2005). Hurricanes <strong>and</strong><br />

global warming. Bulletin of the American Meteorological Society, 86(11):1571–1575.<br />

Pierre, L. S., Kopp, G. A., Surry, D., <strong>and</strong> Ho, T. (2005). The UWO contribution to the<br />

NIST aerodynamic database <strong>for</strong> wind loads on low buildings: Part 2. Comparison of data<br />

with wind load provisions. Journal of <strong>Wind</strong> Engineering <strong>and</strong> Industrial Aerodynamics,<br />

93(1):31–59.<br />

Pinelli, J.-P., Simiu, E., Gurley, K., Subramanian, C., Zhang, L., Cope, A. D., Filliben,<br />

J., <strong>and</strong> Hamid, S. (2004). Hurricane damage prediction model <strong>for</strong> residential structures.<br />

Journal of Structural Engineering, 130(11):1685–1691.<br />

Powell, M., Soukup, G., Cocke, S., Gulati, S., Morisseau-Leroy, N., Hamid, S., Dorst, N.,<br />

<strong>and</strong> Axe, L. (2005). State of florida hurricane loss projection model: Atmospheric science<br />

component. Journal of <strong>Wind</strong> Engineering <strong>and</strong> Industrial Aerodynamics, 93(8):651–674.<br />

Powell, M. D. (1980). Evaluation of diagnostic marine boundary - layer models applied to<br />

tropical cyclones. Monthly Weather Review, 108(6):757–766.<br />

Rosowsky, D. <strong>and</strong> Cheng, N. (1999a). Reliability of light-frame roofs in high-wind regions.<br />

I: <strong>Wind</strong> loads. Journal of Structural Engineering, 125(7):725–733.


BIBLIOGRAPHY 137<br />

Rosowsky, D. <strong>and</strong> Cheng, N. (1999b). Reliability of light-frame roofs in high-wind regions.<br />

II: Reliability analysis. Journal of Structural Engineering, 125(7):734–739.<br />

R.S. Means. (2009). Residential Cost Data. R.S. Means Company.<br />

Sill, B. L. <strong>and</strong> Kozlowski, R. T. (1997). Analysis of storm-damage factors <strong>for</strong> low-rise<br />

structures. Journal of Per<strong>for</strong>mance of Constructed Facilities, 11(4):168–177.<br />

Simiu, E., Changery, J., <strong>and</strong> Filliben, J. J. (1979). Extreme wind speeds at 129 stations in<br />

the contiguous united states. Technical Report NBS Building Science Series 118, National<br />

Institute of St<strong>and</strong>ards <strong>and</strong> Technology.<br />

Simiu, E. <strong>and</strong> Heckert, N. (1996). Extreme wind distribution tails: A ”peaks over threshold”<br />

approach. Journal of Structural Engineering, 122(5):539–547.<br />

Simiu, E., Vickery, P., <strong>and</strong> Kareem, A. (2007). Relation between Saffir-Simpson hurricane<br />

scale wind speeds <strong>and</strong> peak 3-s gust speeds over open terrain. Journal of Structural<br />

Engineering, 133(7):1043–1045.<br />

Skamarock, W., Klemp, J., Dudhia, J., Gill, D., Barker, D., Wang, W., <strong>and</strong> Powers,<br />

J. (2005). A description of the advanced research WRF version 2. Technical report,<br />

Mesoscale <strong>and</strong> Microscale Meteorology Division, National Center <strong>for</strong> Atmospheric Re-<br />

search.<br />

Sparks, P. (2003). <strong>Wind</strong> speeds in tropical cyclones <strong>and</strong> associated insurance losses. Journal<br />

of <strong>Wind</strong> Engineering <strong>and</strong> Industrial Aerodynamics, 91(12-15):1731–1751.<br />

Sparks, P., Schiff, S., <strong>and</strong> Reinhold, T. (1994). <strong>Wind</strong> damage to envelopes of houses <strong>and</strong><br />

consequent insurance losses. Journal of <strong>Wind</strong> Engineering <strong>and</strong> Industrial Aerodynamics,<br />

53(1-2):145–155.


BIBLIOGRAPHY 138<br />

Steenbergen, R., Geurts, C., <strong>and</strong> Bentum, C. V. (2009). Climate change <strong>and</strong> its impact on<br />

structural safety. Heron, 54(1):3–36.<br />

Stewart, M. G., Rosowsky, D. V., <strong>and</strong> Huang, Z. (2003). Hurricane risks <strong>and</strong> economic<br />

viability of strengthened construction. Natural <strong>Hazard</strong>s Review, 4(1):12–19.<br />

Stubbs, N. (1996). Estimation of building damage as a result of hurricanes in the Car-<br />

ribean. Technical report, Organization of American States, General Secretariat, Unit <strong>for</strong><br />

Sustainable Development <strong>and</strong> Environment.<br />

Stubbs, N. <strong>and</strong> Perry, D. C. (1996). Damage simulation model <strong>for</strong> buildings <strong>and</strong> contents<br />

in a hurricane environment. In Ghosh, S. K. <strong>and</strong> Mohammadi, J., editors, Building an<br />

International Community of Structural Engineers, volume 2, pages 989–996, New York,<br />

NY. ASCE.<br />

Tachikawa, M. (1983). Trajectories of flat plates in uni<strong>for</strong>m flow with application to wind-<br />

generated missiles. Journal of <strong>Wind</strong> Engineering & Industrial Aerodynamics, 14(1-3):443–<br />

453.<br />

Twisdale, L., Vickery, P., <strong>and</strong> Steckley, A. (1996). Analysis of hurricane windborne debris<br />

impact risk <strong>for</strong> residential structures. Technical report, Applied Research Associates, Inc.,<br />

Raleigh, NC.<br />

Unanwa, C., McDonald, J., Mehta, K., <strong>and</strong> Smith, D. (2000). The development of wind<br />

damage b<strong>and</strong>s <strong>for</strong> buildings. Journal of <strong>Wind</strong> Engineering <strong>and</strong> Industrial Aerodynamics,<br />

84(1):119–149.<br />

Unanwa, C. O. <strong>and</strong> McDonald, J. R. (2000). Building wind damage prediction <strong>and</strong> mitiga-<br />

tion using damage b<strong>and</strong>s. Natural <strong>Hazard</strong>s Review, 1(4):197–203.<br />

van de Lindt, J. W. <strong>and</strong> Dao, T. N. (2009). Per<strong>for</strong>mance-based wind engineering <strong>for</strong> wood-<br />

frame buildings. Journal of Structural Engineering, 135(2):169–177.


BIBLIOGRAPHY 139<br />

Vanmarcke, E., Lin, N., <strong>and</strong> Yau, S. C. (2010). Quantitative risk analysis of damage to<br />

structures during windstorms: Some multi-scale <strong>and</strong> system-reliability effects. In Li, J.,<br />

Zhao, Y.-G., Chen, J., <strong>and</strong> Peng, Y., editors, Proceedings on the International Symposium<br />

on Reliability Engineering <strong>and</strong> <strong>Risk</strong> <strong>Management</strong>, Shanghai, China. Tongji University<br />

Press.<br />

Vickery, P. <strong>and</strong> Twisdale, L. (1995). <strong>Wind</strong> field <strong>and</strong> filling models <strong>for</strong> hurricane wind speed<br />

predictions. Journal of Structural Engineering, 121(11):1700–1709.<br />

Vickery, P. J., Lin, J., Skerlj, P. F., <strong>and</strong> Twisdale, L. A. (2006a). HAZUS-MH hurricane<br />

model methodology. I: hurricane hazard, terrain, <strong>and</strong> wind load modeling. Natural Haz-<br />

ards Review, 7(2):82–93.<br />

Vickery, P. J., Skerlj, P. F., Lin, J., <strong>and</strong> Twisdale, L. A. (2006b). HAZUS-MH hurricane<br />

model methodology. II: damage <strong>and</strong> loss estimation. Natural <strong>Hazard</strong>s Review, 7(2):94–<br />

103.<br />

Vickery, P. J., Skerlj, P. F., Steckley, A. C., <strong>and</strong> Twisdale, L. A. (2000). Hurricane wind field<br />

model <strong>for</strong> use in hurricane simulations. Journal of Structural Engineering, 126(10):1203–<br />

1221.<br />

Watson, C. C. <strong>and</strong> Johnson, M. E. (2004). Hurricane loss estimation models: Opportunities<br />

<strong>for</strong> improving the state of the art. Bulletin of the American Meteorological Society,<br />

85(11):1713–1726.<br />

Yau, S. C., Lin, N., <strong>and</strong> Vanmarcke, E. (in press). Hurricane damage <strong>and</strong> loss estimation<br />

using an integrated vulnerability model. Natural <strong>Hazard</strong>s Review.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!