Download as pdf or txt
Download as pdf or txt
You are on page 1of 332

Downloaded from ascelibrary.org by 185.106.28.7 on 10/29/20. Copyright ASCE. For personal use only; all rights reserved.

Tsunami Loads and Effects


@Seismicisolation
@Seismicisolation
Tsunami Loads and Effects
Other Titles of Interest
ASCE 7 Hazard Tool. (ASCE, 2016). Delivers a quick, efficient way to look up key design parameters
specified by Standards ASCE/SEI 7-10 and 7-16 through a web-based application that retrieves
load data for each of seven hazards, visualizes them on a map, and generates a unified report of
results. (http://ascetools.online)
Downloaded from ascelibrary.org by 185.106.28.7 on 10/29/20. Copyright ASCE. For personal use only; all rights reserved.

Minimum Design Loads and Associated Criteria for Buildings and Other Structures, ASCE/SEI 7-16.
(ASCE Standard, 2016). Provides requirements for general structural design and includes means
for determining dead, live, soil, flood, wind, snow, rain, atmospheric ice, tsunami, and earth-
quake loads and their combinations that are suitable for inclusion in building codes and other
documents. A detailed commentary of explanatory and supplementary information is included.
(ISBN 978-0-7844-1424-8)

Significant Changes to the Minimum Design Load Provisions of ASCE 7-16, by Gary Chock, P.E.,
D.CE.; S. K. Ghosh, Ph.D.; Michael O’Rourke, Ph.D., P.E.; and T. Eric Stafford, P.E. (ASCE Press,
2018) Describes the revisions to the minimum design load requirements set forth in Standard
ASCE/SEI 7-16. (ISBN 978-0-7844-1457-6)

Wind Loads: Guide to the Wind Load Provisions of ASCE  7-16, by William L. Coulbourne, P.E.,
and T. Eric Stafford, P.E. (ASCE Press, 2020) Sets forth a detailed and authoritative interpretation
of the wind load provisions of Minimum Design Loads and Associated Criteria for Buildings and
Other Structures, Standard ASCE/SEI 7-16. (ISBN 978-0-7844-1526-9)

Snow Loads: Guide to the Snow Load Provision of ASCE  7-16, by Michael O’Rourke, Ph.D., P.E.
(ASCE Press, 2017) Supplies a detailed, authoritative explanation of the snow load provisions con-
tained in Minimum Design Loads and Associated Criteria for Buildings and Other Structures, Stan-
dard ASCE/SEI 7-16. (ISBN 978-0-7844-1456-9)

Seismic Loads: Guide to the Seismic Load Provisions of ASCE 7-16, by Finley A. Charney, Ph.D.,
P.E.; Thomas F. Heausler, P.E., S.E.; and Justin D. Marshall, Ph.D., P.E. (ASCE Press, 2020) Pro-
vides clear, authoritative explanations of the seismic design provisions contained in Minimum
Design Loads and Associated Criteria for Buildings and Other Structures, Standard ASCE/SEI 7-16
(ISBN 978-0-7844-1550-4)

Rain Loads: Guide to the Rain Load Provisions of ASCE 7-16, by Michael O’Rourke, Ph.D., P.E.,
and Aaron R. Lewis, P.E. (ASCE Press, 2020). Provides a comprehensive overview of the rain load
provisions in Minimum Design Loads and Associated Criteria for Buildings and Other Structures,
Standard ASCE/SEI 7-16. (ISBN 978-0-7844-4155-3)

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Tsunami Loads and Effects
Downloaded from ascelibrary.org by 185.106.28.7 on 10/29/20. Copyright ASCE. For personal use only; all rights reserved.

Guide to the Tsunami Design


Provisions of ASCE 7-16

Ian N. Robertson, Ph.D., S.E.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Library of Congress Cataloging-in-Publication Data

Names: Robertson, Ian N. (Ian Nicol), author.


Downloaded from ascelibrary.org by 185.106.28.7 on 10/29/20. Copyright ASCE. For personal use only; all rights reserved.

Title: Tsunami loads and effects : guide to the tsunami design provisions of ASCE 7-16 / Ian N. Robertson, Ph.D., S.E.
Description: Reston, Virginia : American Society of Civil Engineers, [2020] |
Includes bibliographical references and index.
Identifiers: LCCN 2017057404| ISBN 9780784414972 (soft cover : alk. paper) | ISBN 9780784480212 (pdf) |
ISBN 9780784480854 (epub)
Subjects: LCSH: Tsunami resistant design--Standards--United States. | Tsunami damage--Prevention. | Ocean
waves.
Classification: LCC TA654.55 .R63 2018 | DDC 624.1/76--dc23
LC record available at https://lccn.loc.gov/2017057404

Published by American Society of Civil Engineers


1801 Alexander Bell Drive
Reston, Virginia 20191-4382
www.asce.org/publications | ascelibrary.org

Any statements expressed in these materials are those of the individual authors and do not necessarily repre-
sent the views of ASCE, which takes no responsibility for any statement made herein. No reference made in this
publication to any specific method, product, process, or service constitutes or implies an endorsement, recom-
mendation, or warranty thereof by ASCE. The materials are for general information only and do not represent a
standard of ASCE, nor are they intended as a reference in purchase specifications, contracts, regulations, stat-
utes, or any other legal document. ASCE makes no representation or warranty of any kind, whether express or
implied, concerning the accuracy, completeness, suitability, or utility of any information, apparatus, product, or
process discussed in this publication, and assumes no liability therefor. The information contained in these mate-
rials should not be used without first securing competent advice with respect to its suitability for any general or
specific application. Anyone utilizing such information assumes all liability arising from such use, including but
not limited to infringement of any patent or patents.

ASCE and American Society of Civil Engineers—Registered in US Patent and Trademark Office.
Photocopies and permissions. Permission to photocopy or reproduce material from ASCE publications can be
requested by sending an email to permissions@asce.org or by locating a title in the ASCE Library (https://asceli-
brary.org) and using the “Permissions” link.

Errata: Errata, if any, can be found at https://doi.org/10.1061/9780784414972.

Copyright © 2020 by the American Society of Civil Engineers.


All Rights Reserved.

ISBN 978-0-7844-1497-2 (print)


ISBN 978-0-7844-8117-2 (PDF)
Manufactured in the United States of America.

26 25 24 23 22 21 20     1  2  3  4  5

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Contents

Foreword, by Gary Chock, S.E., F.SEI, Dist. M. ASCE  ix


Preface xiii
Overview of Contents  xv
Downloaded from ascelibrary.org by 185.106.28.7 on 10/29/20. Copyright ASCE. For personal use only; all rights reserved.

Unit Conversions  xvii

Chapter 1. Introduction 1
1.1  Past Tsunamis  1
1.2  Tsunami Generation Mechanisms  12
1.3  Warning and Evacuation  14
1.4  Development of ASCE 7-16, Chapter 6  17

Chapter 2. Prototype Buildings  21


2.1  Reinforced Concrete Moment-Resisting Frame Office Building  23
2.2  Reinforced Concrete Shear Wall Residential Building  24
2.3  Seismic and Wind Design  25

Chapter 3. General Requirements  27


3.1 Scope  30
3.2  Probabilistic Tsunami Hazard Analysis  32
3.3  Tsunami Design Zone Maps  32
3.4  Definitions, Symbols, and Notation  34

Chapter 4. Tsunami Risk Categories  37

Chapter 5. Tsunami Flow Characteristics  39


5.1  TRC II and III Buildings and Other Structures  39
5.2  TRC IV Buildings and Other Structures  41
5.3  Sea Level Change  42

Chapter 6. Inundation Depths and Flow Velocities Based on Runup  43


6.1  Maximum Inundation Depth and Flow Velocities  44
6.2  Energy Grade Line Analysis  46
6.3  Terrain Roughness  58
6.4  Tsunami Bores  61
6.5  Amplified Flow Velocities  62

v
@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


vi Contents

Chapter 7. Site-Specific Tsunami Inundation  63


7.1  Tsunami Waveform  65
7.2  Tsunamigenic Sources  67
7.3  Earthquake Rupture Unit Source Tsunami Functions  67
7.4  Treatment of Modeling and Natural Uncertainties  68
7.5  Offshore Tsunami Amplitude  68
7.6  Procedures for Determining Tsunami Inundation and Runup  68

Chapter 8. Structural Design Procedures for Tsunami Effects  71


Downloaded from ascelibrary.org by 185.106.28.7 on 10/29/20. Copyright ASCE. For personal use only; all rights reserved.

8.1  Performance of Tsunami Risk Category II and III Buildings and Other
Structures 72
8.2  Performance of Tsunami Risk Category III Critical Buildings and TRC IV
Buildings and Other Structures  73
8.3  Structural Performance Evaluation  73
8.4  Minimum Fluid Density for Tsunami Loads  81
8.5  Flow Velocity Amplification  82
8.6  Directionality of Flow  85
8.7  Minimum Closure Ratio for Load Determination  86
8.8  Minimum Number of Tsunami Flow Cycles  90
8.9  Seismic Effects on the Foundations Preceding Local Subduction Zone
Maximum Considered Tsunami  90
8.10  Physical Modeling of Tsunami Flow, Loads, and Effects  91

Chapter 9. Hydrostatic Loads  93


9.1  Buoyancy  93
9.2  Unbalanced Lateral Hydrostatic Force  105
9.3  Residual Water Surcharge Load on Floors and Perimeter Walls  108
9.4  Hydrostatic Surcharge Pressure on Foundations  110

Chapter 10. Hydrodynamic Loads  111


10.1  Simplified Equivalent Uniform Lateral Static Force  112
10.2  Detailed Hydrodynamic Lateral Forces  113
10.3  Hydrodynamic Pressures Associated with Slabs  133

Chapter 11. Debris Impact Loads  157


11.1  Alternative Simplified Debris Impact Static Load  160
11.2  Wood Poles and Logs  161
11.3  Debris Impacts by Floating Vehicles  164
11.4  Debris Impacts of Submerged Tumbling Boulder and Concrete
Debris 167
11.5  Site Hazard Assessment for Shipping Containers, Ships, and Barges  167

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Contents vii

11.6  Shipping Containers  171


11.7  Extraordinary Debris Impacts  173
11.8  Alternative Methods of Response Analysis  174
11.9  Detailed Debris Impact Calculation for Monterey, California, Office
Building  175

Chapter 12. Foundation Design  177


12.1  Resistance Factors for Foundation Stability Analyses  179
12.2  Load and Effect Characterization  180
Downloaded from ascelibrary.org by 185.106.28.7 on 10/29/20. Copyright ASCE. For personal use only; all rights reserved.

12.3  Alternative Foundation Performance-Based Design Criteria  199


12.4  Foundation Countermeasures  199

Chapter 13. Structural Countermeasures  203


13.1  Open Structures  203
13.2  Tsunami Barriers  204

Chapter 14. Tsunami Vertical Evacuation Refuge Structures  207


14.1  Minimum Inundation Elevation and Depth  210
14.2  Refuge Live Load  212
14.3  Laydown Impacts  214
14.4  Information on Construction Documents  214
14.5  Peer Review  214

Chapter 15. Designated Nonstructural Components and Systems  217


15.1  Performance Requirements  217

Chapter 16. Tsunami Risk Categories III and IV Non-Building Structures  221
16.1  Requirements for TRC III Non-Building Structures  221
16.2  Requirements for TRC IV Non-Building Structures  221

Chapter 17. Frequently Asked Questions  223

References 231
Appendix. Seaside, Oregon: Full Design Example  239
Index 313
About the Author  321

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Downloaded from ascelibrary.org by 185.106.28.7 on 10/29/20. Copyright ASCE. For personal use only; all rights reserved.

Tsunami Loads and Effects


@Seismicisolation
@Seismicisolation
Foreword
Downloaded from ascelibrary.org by 185.106.28.7 on 10/29/20. Copyright ASCE. For personal use only; all rights reserved.

The design of a structure to resist tsunami loads and effects should be undertaken with the
most serious intent because of the very high mortality rate of tsunami events. In high tsunami
hazard regions, tsunami loading on structural components can be extreme, and inundation
depths can amount to tens of feet of water flowing at high velocity. As we have witnessed in
past tsunamis, under such conditions people unable to leave the tsunami inundation zone
almost certainly become fatalities unless they can seek refuge in robust structures of suf-
ficient height. Therefore, when engaged in the design of any building that falls within the
scope of ASCE 7-16, Chapter 6, engineers should bear in mind at all times that their work is to
create the sanctuary that may be the only means of survival against otherwise nearly impos-
sible odds. It is recommended that the beginning of every designer’s involvement include
meeting the community in person to emphasize this relationship. Furthermore, the design
professionals should also give serious study of past events, such as the 2004 Indian Ocean
tsunami and the 2011 Tohoku Japan tsunami, to gain a further understanding of the effects
of tsunamis on structures; that will serve as visualization when working through the design
requirements intended to defend against the many mechanisms of possible failure.
The ASCE 7 Tsunami Loads and Effects Subcommittee (TLESC) that was charged in Feb-
ruary 2011 to develop the design requirements included experts that comprised the diversity
of knowledge of seismology, probabilistic tsunami hazard analysis and inundation modeling,
coastal, hydraulic, structural, and geotechnical engineering, and structural reliability analy-
sis. The ASCE 7 tsunami design provisions are the timely culmination of more than a decade
of recent engineering research, including that of the author, Ian Robertson of the University
of Hawai‘i, which was particularly aimed to produce practical design applications for struc-
tural loadings.
This multi-disciplinary insight was brought to bear on the reconnaissance of the March
11, 2011, Great East Japan Earthquake and Tsunami, where detailed investigations of struc-
tures were targeted based on remote sensing and nationwide damage surveys conducted
immediately after the event by 300 Japanese researchers coordinated as the 2011 Tohoku
Earthquake Tsunami Joint Survey Group. The ASCE tsunami reconnaissance began in April of

ix
@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


x Foreword
Downloaded from ascelibrary.org by 185.106.28.7 on 10/29/20. Copyright ASCE. For personal use only; all rights reserved.

2011 in collaboration with tsunami researchers mostly affiliated with Dr. Tomoya Shibayama
of Waseda University and the Japan Society of Civil Engineers, Kyoto University, the Tokyo
Institute of Technology, the University of Tokyo, Yokohama National University, and Saitama
University. The ASCE tsunami reconnaissance team performed a series of investigations that
included field measurements and material samples sufficient to comprise independent val-
idation case examples of structural failure mechanisms for the anticipated tsunami design
formulations. A second group from the University of Hawai‘i and Oregon State University,
funded by a National Science Foundation RAPID grant, used ground-based LiDAR capabili-
ties to collect detailed three-dimensional (3D) structural data for specific structures, which
were then used as input for the structural nonlinear analysis modeling that was compared
against the actual structural damage and deformations.
In combination with past reconnaissance from 2005 onward conducted by members
of the ASCE TLESC, the ASCE 7 tsunami design provisions were subjected to a series of
structural analysis case studies together with numerical modeling and reliability analysis for
validation prior to being presented for ultimate approval by the ASCE Standards Committee
accredited consensus process. Subsequently, it was approved for incorporation into the 2018
International Building Code that is used as the model code for all states in the United States.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Foreword xi

Presently, the tsunami design requirements apply to the five western states of Alaska, Wash-
ington, Oregon, California, and Hawai‘i.
The reader will find this book to be an excellent companion to ASCE  7-16, Minimum
Design Loads and Associated Criteria for Buildings and Other Structures. It is intended to
provide guidance to structural engineers and other design professionals charged with an
actual project. The ASCE 7-16 standard includes an extensive commentary to the provisions
of Chapter 6, “Tsunami Loads and Effects.” Nevertheless, Ian Robertson has composed much
needed additional discusion and design examples to explain the requirements. It would be an
Downloaded from ascelibrary.org by 185.106.28.7 on 10/29/20. Copyright ASCE. For personal use only; all rights reserved.

effective practice to refer to the ASCE 7-16 Commentary while reading this book to garner a
comprehensive context to the requirements of the provisions.
As chair of the ASCE 7 Tsunami Loads and Effects Subcommittee, I commend the dedi-
cated work of its members that resulted in the first national, consensus-based standard for
tsunami-resilient design, and I particularly recommend this work of Ian Robertson that pro-
vides a truly useful design guide for practitioners engaged in the defense of our vital coastal
communities and critical facilities.

Gary Chock, S.E., F.SEI, Dist. M. ASCE


Structural Engineer, Hawai‘i and California
ASCE Structural Engineering Institute Fellow
Distinguished Member ASCE
Diplomate, Coastal Engineer, of the Academy of Coastal, Ocean, Port and Navigation Engineers
ASCE 7 Tsunami Loads and Effects Subcommittee Chair

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Downloaded from ascelibrary.org by 185.106.28.7 on 10/29/20. Copyright ASCE. For personal use only; all rights reserved.

Tsunami Loads and Effects


@Seismicisolation
@Seismicisolation
Preface
Downloaded from ascelibrary.org by 185.106.28.7 on 10/29/20. Copyright ASCE. For personal use only; all rights reserved.

Although there have been devastating tsunamis throughout recorded history, it was the
Indian Ocean tsunami of December 26, 2004, that raised the profile of tsunami research in
the general population around the world. For decades, Japan had been designing and con-
structing coastal defensive structures to resist tsunami loads because of their long and pain-
ful history of loss of life and property due to tsunami inundation. Researchers around the
world had also been studying the generation and open ocean propagation of tsunamis with
the intent of improving warning systems and public awareness. However, it took the enor-
mous loss of life and infrastructure damage caused around the Indian Ocean for other coun-
tries to take note and begin to address the tsunami hazard in earnest.
In the United States, Honolulu City and County was the only jurisdiction that required
any consideration of tsunami loads and effects in the design of coastal buildings (Honolulu
City & County 1984). These provisions were rather rudimentary and were based on research
performed in the 1970s (Dames & Moore 1980). These provisions also lacked any tsunami haz-
ard maps, so there were very few actual applications of the provisions. A workshop held in
Tacoma, Washington, in November 2002 was one of the first attempts to include structural
engineering in tsunami research in the United States (Walsh et al. 2002). A preliminary study
stemming from that workshop came to the conclusion that properly engineered buildings
and other structures could survive the effects of a major tsunami with damage only to non-
structural elements below the flow level (Yeh et al., 2005).
In 2004, FEMA funded a project through the Applied Technology Council of Redwood
City, California, to develop Guidelines for Design of Structures for Vertical Evacuation from
Tsunamis. First published as FEMA P-646 in 2008, this document was updated subsequent to
the Tohoku tsunami of March 11, 2011, with the second edition published in 2012 (FEMA 2012).
These guidelines were written specifically for the design of designated vertical evacuation
structures but made no attempt to address the general building stock. As such, the provi-
sions could deliberately err on the conservative side without jeopardizing the cost of coastal
construction in general. This document was not written in mandatory language, so it was not
amenable to implementation in legally binding building codes.
Following the 2004 Indian Ocean tsunami, considerable laboratory experimentation and
computational research has been funded in the United States and elsewhere. Field surveys of

xiii
@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


xiv Preface

damaged and surviving structures after the Indian Ocean tsunami and more recent Samoa tsu-
nami (2009) and Chile tsunami (2010) provided additional insight into fluid-structure interac-
tion during coastal inundation resulting from tsunamis. Based on the results of this laboratory
research and the post-tsunami field investigations, a group of 30 researchers and engineers col-
laborated to form a new ASCE 7 Tsunami Loads and Effects Subcommittee (TLESC). In February
2011, this group, under the leadership of Gary Chock, was formally accepted as a fully functioning
subcommittee of ASCE 7. At the time, Standard ASCE/SEI 7-16, Minimum Design Loads and Asso-
ciated Criteria for Buildings and Other Structures, considered seismic, wind, rain, flood, snow,
Downloaded from ascelibrary.org by 185.106.28.7 on 10/29/20. Copyright ASCE. For personal use only; all rights reserved.

live, dead, and other loads, but had no provisions requiring design for tsunami loads.
Only one month later, the Great East Japan Earthquake with a moment magnitude of 9.2
generated a tsunami that devastated many communities along the northeast (Tōhoku) coast of
Honshu Island, commonly referred to as the Tohoku tsunami. ASCE funded seven members of
TLESC to travel to the affected areas one month after the tsunami to gather as much informa-
tion as possible to aid in development of the design provisions. Analysis of buildings and other
structures damaged during the Tohoku tsunami contributed significantly to the subcommit-
tee’s work on the tsunami design provisions (Chock et al. 2013a, b; Carden et al. 2015).
The TLESC worked for three and a half years to develop a draft chapter on tsunami loads
and effects for incorporation in ASCE 7-16. After multiple internal subcommittee ballots, this draft
chapter was submitted to the ASCE 7 Main Committee for its first ballot in July 2014. After a total
of eight main committee ballots and a period for public comment, the consensus-based docu-
ment of requirements and accompanying commentary was accepted into ASCE 7-16 as Chapter 6,
“Tsunami Loads and Effects,” on March 11, 2016, exactly five years after the Tohoku tsunami.
This guide serves as a companion document to the new ASCE 7-16, Chapter 6 on “Tsu-
nami Loads and Effects.” Since this is the first exposure to tsunami design in the United
States for the majority of structural and geotechnical engineers, architects, building officials,
contractors, and others involved in the building design and construction process, this com-
panion manual was developed to guide new users through the application of tsunami design
to coastal buildings and other structures. The target audience for this book is therefore
structural and geotechnical engineers, architects, building officials, and others who cur-
rently use ASCE 7. It is my intent to help the reader understand the background to the code
provisions while also providing clear and detailed examples demonstrating the application of
the provisions to typical coastal buildings.
The examples presented in the guide often refer to sections, equations, tables, and
figures in ASCE  7-16. All such items are referred to directly, without specific reference to
ASCE 7-16. For instance, a specific example might contain the statement, “The illustration is
provided in Table 6.12-1 of the standard.”
References to sections, equations, tables, and figures that are unique to the guide are
always preceded by the letter G and use bold text. For example, the text may state that the
distribution of forces along the height of the structure is provided in Table G12-3 and illus-
trated in Figure G12-5. In this citation, the number 12 is the guide chapter number, and the
number after the dash is the sequence number of the item (i.e., third table or fifth figure).

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Overview of Contents
Downloaded from ascelibrary.org by 185.106.28.7 on 10/29/20. Copyright ASCE. For personal use only; all rights reserved.

This guide describes the basis for all provisions in the new ASCE 7-16 Chapter 6 on “Tsu-
nami Loads and Effects” and demonstrates by example how each of the provisions applies
to prototypical coastal buildings. Chapter 1 of this guide provides an overview of the back-
ground for development of the tsunami design provisions in ASCE 7-16. It describes many of
the observations made during past tsunamis that contributed to development of particular
design provisions, as well as a brief history of the development of tsunami design in the
United States. Chapter 2 introduces the prototypical buildings that are used throughout the
rest of the manual as example applications of the design provisions. Chapter 3 covers the
first three sections of ASCE 7-16 Chapter 6, “General Requirements, Definitions, and Symbols
and Notation.” The remaining chapters of this manual are numbered to correspond to the
ASCE 7-16 Chapter 6 subsections, as follows:

Chapter 4 – ASCE 7-16 Section 6.4 Tsunami Risk Categories


Chapter 5 – ASCE 7-16 Section 6.5 Analysis of Design Inundation Depth and Flow Velocity
Chapter 6 – ASCE 7-16 Section 6.6 Inundation Depths and Flow Velocities Based on Runup
Chapter 7 – ASCE 7-16 Section 6.7 Inundation Depths and Flow Velocities Based on Site-
Specific Probabilistic Tsunami Hazard Analysis
Chapter 8 – ASCE 7-16 Section 6.8 Structural Design Procedures for Tsunami Effects
Chapter 9 – ASCE 7-16 Section 6.9 Hydrostatic Loads
Chapter 10 – ASCE 7-16 Section 6.10 Hydrodynamic Loads
Chapter 11 – ASCE 7-16 Section 6.11 Debris Impact Loads
Chapter 12 – ASCE 7-16 Section 6.12 Foundation Design
Chapter 13 – ASCE 7-16 Section 6.13 Structural Countermeasures for Tsunami Loading
Chapter 14 – ASCE 7-16 Section 6.14 Tsunami Vertical Evacuation Refuge Structures
Chapter 15 – ASCE 7-16 Section 6.15 Designated Nonstructural Components and Systems
Chapter 16 – ASCE 7-16 Section 6.16 Non-Building Tsunami Risk Category III and IV Structures

xv
@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


xvi Overview of Contents

It is assumed that the reader has a copy of ASCE 7-16


available for reference while using this guide. References Read the
Commentary!
to sections and equations in ASCE  7-16 are prefaced with
“ASCE 7,” whereas references to the ASCE 7-16 Commentary The reader is strongly
are prefaced with “ASCE 7 Commentary.” References without encouraged to read the
ASCE 7-16 Commentary
these prefaces are to sections within this guide.
relating to each code
Because the ASCE  7-16 Chapter 6 “Tsunami Loads and provision to gain a
Effects” provisions are new and likely to be unfamiliar to the better understanding of
Downloaded from ascelibrary.org by 185.106.28.7 on 10/29/20. Copyright ASCE. For personal use only; all rights reserved.

user, extra effort was made by the subcommittee to develop the code requirements.
a detailed commentary. The reader is strongly encouraged to
read the commentary relating to each code provision to gain
a better understanding of the code requirements.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Unit Conversions
Downloaded from ascelibrary.org by 185.106.28.7 on 10/29/20. Copyright ASCE. For personal use only; all rights reserved.

US customary units International System of Units (SI)


1 inch (in.) 25.4 millimeters (mm)

1 foot (ft) 0.3048 meter (m)

1 statute mile (mi) 1.6093 kilometers (km)

1 square foot (ft2) 0.0929 square meter (m2)

1 cubic foot (ft3) 0.0283 cubic meter (m3)

1 foot per second (ft/s) 0.3048 meter per second (m/s)

1 slug (sl) (mass) 0.4536 kilogram (kg)

1 pound (lb) (force) 4.4482 newtons (N)

1 pound per lineal foot (lb/ft, plf) 0.0146 kilonewtons per lineal meter (kN/m)

1 pound per square foot (lb/ft2, psf) 0.0479 kilonewton per square meter (kN/m2, kPa)

1 pound per cubic foot (lb/ft3) 0.157 kilonewtons per cubic meter (kN/m3)

1 slug per cubic foot (sl/ft3) 512.6 kilograms per cubic meter (kg/m3)

xvii
@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Downloaded from ascelibrary.org by 185.106.28.7 on 10/29/20. Copyright ASCE. For personal use only; all rights reserved.

Tsunami Loads and Effects


@Seismicisolation
@Seismicisolation
1
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 09/20/20. Copyright ASCE. For personal use only; all rights reserved.

Introduction

1.1  Past Tsunamis


Trans-oceanic tsunamis are generated regularly by large subduction zone earthquakes. Over
the last 70 years (1950–2020), 24 trans-oceanic tsunamis have occurred around the world but
primarily in the Pacific Ocean (Wikipedia 2020b). All these tsunamis resulted in coastal inun-
dation locally near the source with associated loss of life and property. Many were also large
enough to result in inundation on distant shores. Most notably, the Indian Ocean tsunami
caused by a 9.1-magnitude earthquake off the west coast of northern Sumatra resulted in more
than 225,000 deaths in countries bordering the Indian Ocean (Wikipedia 2020c). This was a
wake-up call that led to development of tsunami warning systems throughout the Pacific and
Indian Oceans, as well as the Caribbean and other tsunami-prone coastlines. Subsequent dam-
aging tsunamis affecting Samoa in 2009 (189 fatalities), Chile in 2010 (507 fatalities and $15 to
$30 billion in losses), and Tohoku Japan in 2011 (more than 18,000 fatalities and approximately
$360 billion in losses) (Wikipedia 2020b) reinforced the need for improved tsunami prepared-
ness both in terms of warning systems and evacuation planning and in the form of enhanced
resilience of coastal communities. By improving the performance of critical and essential facil-
ities and all buildings that are taller than the anticipated flow depth, a coastal community can
reduce the number of deaths and casualties, limit the extent of structural damage, reduce the
financial losses, and facilitate more rapid recovery after a damaging tsunami.
Post-tsunami reconnaissance has gathered valuable information regarding the perfor-
mance of engineered and nonengineered structures when subjected to high-velocity tsu-
nami flow. In particular, the Tohoku tsunami demonstrated the potential devastating effects
of a major tsunami on coastal communities and their economies.
During past tsunamis, most timber-framed residential structures that experienced flow
depths more than 8 ft either floated or were destroyed (Figures G1-1, G1-2, and G1-3). Unre-
inforced and lightly reinforced masonry structures commonly experienced out-of-plane wall
failures (Figure G1-4). Some reinforced concrete, reinforced masonry, and structural steel-
framed buildings were also destroyed during the Tohoku tsunami (Figures G1-5 and G1-6).

1
@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


2 Tsunami Loads and Effects
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 09/20/20. Copyright ASCE. For personal use only; all rights reserved.

Figure G1-1. Timber-framed residential structure in Noda, Japan, that floated intact off its foundation
(left) (N 40.103443; E 141.822927), or was completely destroyed by the tsunami flow (right) (N 40.103665;
E 141.82434).

Figure G1-2. Surviving 3-story concrete wall structure in Minamisanriku, Japan, covered by timber
from failed residential buildings (N 38.6804; E 141.4416).

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Introduction 3
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 09/20/20. Copyright ASCE. For personal use only; all rights reserved.

Figure G1-3. Overturned timber-framed single-family residence, Minamisanriku, Japan, (N 38.68107;


E 141.44052).

Figure G1-4. Out-of-plane failure of unreinforced masonry walls during the Palu tsunami: (left) S 0.886;
E 119.8557; (right) S 0.8848; E 119.8547.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


4 Tsunami Loads and Effects
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 09/20/20. Copyright ASCE. For personal use only; all rights reserved.

Figure G1-5. Failure of structural steel buildings in Otsuchi, Japan,: (left) N 39.360323; E 141.905965; and
(right) N 39.360788; E 141.905606.
Source: Right side photo courtesy of Gary Chock.

Figure G1-6. (Left) Overturned reinforced concrete building in Onagawa, Japan, (N 38.440723; E
141.446484); and (right) failed reinforced concrete column.

However, many reinforced concrete and structural steel buildings survived structurally,
even though flow depths often exceeded the roof elevation (Figures G1-7 and G1-8). The
Matsubara Community apartment building shown in Figure G1-7 was constructed in 2007 as
a vertical evacuation building to provide a refuge location at the roof level. Even though the
flow overtopped the roof by 2 ft (0.7 m), 44 refugees survived on the roof because the building
was able to resist the tsunami loads (Fraser et al. 2012). The deep-piled foundation system was
able to support the structure even with as much as 6 ft (2 m) scour around the corners of the
building (EEFIT 2011). Figure G1-8 shows the two Marine Pal buildings on the shorefront in
Onagawa, Japan, that were almost completely overtopped but survived with only nonstruc-
tural damage. Survivor video from the top of one of these buildings provides a vivid illustra-
tion of the power of the tsunami flow while also demonstrating that well-designed buildings

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Introduction 5
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 09/20/20. Copyright ASCE. For personal use only; all rights reserved.

Figure G1-7. (Left) Four-story Matsubara Community apartment building in Minamisanriku, Japan, built
in 2007 as a designated vertical evacuation building, survived even though the flow overtopped the roof
by 2 ft (0.7 m) (N 38.67366; E 141.44537); (right) during the tsunami, 44 people survived by seeking refuge
on the roof evacuation area.

Figure G1-8. Marine Pal building complex on the dock in Onagawa, Japan (N 38.44188; E 141.44718). These
reinforced concrete buildings survived even though the flow exceeded all but the very highest turret
roof levels, saving some of those who sought refuge on the roof.

can survive these loads with only nonstructural damage. Figure G1-9 shows a 4-story desig-
nated vertical evacuation refuge building in Kesennuma, Japan, where 200 people survived
flow depths up to the third-floor level. An adjacent 5-story building, which was not desig-
nated for vertical evacuation, was nevertheless used by 120 refugees who survived on the
upper levels and roof (Fraser et al. 2012). During the Tohoku tsunami, numerous designated
vertical evacuation buildings were used, saving at least 5,000 people caught in the inunda-
tion zone (Fraser et al. 2012). In addition, it is estimated that tens of thousands of people were
saved by seeking refuge in other buildings not designated or designed for tsunami loads. In
the town of Ishinomaki, Japan alone, Fraser et al. (2012) estimate that “About 260 official and
unofficial evacuation places were used in total, providing refuge to around 50,000 people.”

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


6 Tsunami Loads and Effects
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 09/20/20. Copyright ASCE. For personal use only; all rights reserved.

Figure G1-9. (Left) Designated vertical evacuation building in Kesennuma, Japan, that was used by 200
people to survive the tsunami (N 38.8916; E 141.5876); (right) adjacent 5-story building that was not
designated as a vertical evacuation site, but nevertheless served as a refuge for 120 people during the
tsunami (N 38.8911; E 141.5886).

Unfortunately, a number of designated vertical evacuation structures were not high


enough to protect those who used them as a refuge. Figure G1-10 shows the City Hall Com-
munity Center and Gymnasium in Rikuzentakata, Japan, where the second floor was an
officially designated tsunami refuge. However, the waves swept through the refuge level
resulting in at least seven deaths (Chock et al. 2013b). Figure G1-11 shows the steel-framed
Emergency Management Center in Minamisanriku, Japan, where 21 emergency responders
lost their lives when the flow overtopped the designated refuge location on the roof of the
building (Chock et al. 2013b). Particular care was taken while developing the ASCE 7-16 (2017a)
tsunami provisions to ensure that tsunami vertical evacuation refuge structures would pro-
vide refuge areas well above the 2,500-year probabilistically determined flow depth so as to

Figure G1-10. The second floor of the Rikuzentakata City Hall Community Center and Gymnasium was
designated as an official tsunami vertical evacuation center but was completely inundated leading to
loss of life for almost all evacuees who sought refuge in the building (N 39.0139; E 141.6339).

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Introduction 7
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 09/20/20. Copyright ASCE. For personal use only; all rights reserved.

Figure G1-11. Minamisanriku Emergency Management Center: (a) before the tsunami with 30 occupants; (b)
with 24 refugees on the roof-level designated tsunami refuge; (c) which was overtopped leading to the loss of
15 lives; (d) leaving only nine of the original building occupants alive after the tsunami (N 38.6778; E 141.4464).
Source: Chock et al. (2013b).

avoid repeating these mistakes. Applying the design procedures in ASCE 7-16 (2017a) will also
result in a vertical evacuation refuge structure with less than 1% probability of structural
failure even when subjected to the 2,500-year maximum considered tsunami (MCT) (Chock
et al. 2016).
Figure G1-12 shows an interesting comparison between two 3-story buildings that
structurally survived the same flow conditions, but with significantly different levels of non-
structural damage. Because the reinforced concrete building consisted primarily of rein-
forced concrete walls, there was limited damage to the building, whereas the open frame
structure in the structural steel building meant that virtually all nonstructural elements were
stripped away by the flow, leaving only the bare steel frame (Figure G1-12).

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


8 Tsunami Loads and Effects
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 09/20/20. Copyright ASCE. For personal use only; all rights reserved.

Figure G1-12. Comparative building resilience; reinforced concrete wall structure and steel frame structure
located side by side in Otsuchi, Japan: (left) rear view; (right) front view facing the ocean (N 39.35698;
E 141.90697).
Source: (Left) courtesy of Ioan Nistor; (right) courtesy of David Kriebel.

A large percentage of taller reinforced concrete and structural steel buildings survived
with only nonstructural damage at their lower levels (Figure G1-13). The ASCE tsunami survey
team and their Japanese collaborators are not aware of any 6-story and higher reinforced
concrete or structural steel building that collapsed during the Tohoku tsunami. It should be
noted that Japanese seismic design, which would have been required for all buildings along the
Tohoku coastline, is somewhat more conservative than US seismic design, relying less on the
nonlinear response of the structure. Nevertheless, it is anticipated that modern taller buildings
along the US Pacific coastline should also perform well structurally during future tsunamis.

Figure G1-13. (Left) Six-story steel-framed hotel in Taro, Japan, that survived flow to the third-floor
level and has since been repaired (N 39.7378; E 141.9757); (right) 7-story reinforced concrete hotel in
Rikuzentaka, Japan, that survived flow intense enough to cause a reinforced concrete wall blowout at
the first level, but has since been demolished (N 39.009; E 141.6314).

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Introduction 9

Figure G1-14 shows a pair of 5-story buildings in Rikuzentakata, Japan, that experienced a
maximum flow depth 3 ft over the fifth-floor level. These buildings are oriented parallel to the
shoreline, so they experienced the hydrodynamic loads on their broad face. They are reinforced
concrete buildings with shear walls providing lateral load resistance. These buildings are similar
in layout and structural system to the residential prototype buildings presented in Chapter 2.
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 09/20/20. Copyright ASCE. For personal use only; all rights reserved.

Figure G1-14. Five-story reinforced concrete residential buildings in Rikuzentakata, Japan,


(N 39.0099; E 141.6443) that survived inundation to the fifth level.

Numerous ports and harbors were severely damaged during the Tohoku tsunami. Ware-
houses and storage buildings often suffered significant structural damage (Figure G1-15), and
fuel storage tanks suffered buckling from hydrostatic and/or hydrodynamic loads or debris
impact, or were washed off their foundations and became floating debris, often punctur-
ing and spilling their contents (Figure G1-16). Wharves and piers performed well during the
Tohoku tsunami with damage limited to uplifting of the access panels or grating covering the
gap between the pile-supported deck and the adjacent slab-on-grade apron (Figure G1-17).
These access panels appear to have served a pressure-relieving function, thus reducing the
maximum pressure below the elevated deck and protecting them from uplift damage. During
the Indian Ocean tsunami, an elevated wharf structure in Khao Lak Harbor, Thailand, suf-
fered extensive damage from uplift of the precast deck slabs [Figure G1-18(a)]. Similar dam-
age was observed in a building in Patong Beach during the same tsunami [Figure G1-18(b)].
Although bridge design is not covered by ASCE 7-16, it would seem incomplete not to
mention the extensive damage to road and railway bridges that has occurred during past
tsunamis. Figure G1-19 shows a highway bridge in Rikuzentakata, Japan, and a railway bridge
in Otsuchi, Japan, where the deck and girders were washed off the pier supports. Pier fail-
ure was also noted in a number of bridge failures (Figure G1-20), as well as significant scour
damage to abutments at either end of the bridge [Figure G1-19(a)] (Chock et al. 2013b).
@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


10 Tsunami Loads and Effects
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 09/20/20. Copyright ASCE. For personal use only; all rights reserved.

Figure G1-15. (a) Reinforced concrete warehouse in Onagawa, Japan, damaged by the Tohoku tsunami
(N 38.4458; E 141.4481); (b) structural steel warehouse in Yuriage, Japan, damaged by the Tohoku
tsunami (N 38.1688; E 170.9509).

Figure G1-16. (a) Damage to fuel storage tanks in Kuji Port during the Tohoku tsunami (N 40.1960;
E 141.7962); (b) fuel storage tank in Kesennuma, Japan, that floated and ruptured owing to impact
damage (N 38.8895; E 141.5917).

Figure G1-17. (a) Access panel uplift at connection between pile-supported pier on right and slab-on-
grade apron on left in Ofunato, Japan (N 39.0649; E 141.7245); (b) pressure-relieving grating washed
away by tsunami in Taro, Japan (N 39.7356; E 141.9770).
@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Introduction 11
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 09/20/20. Copyright ASCE. For personal use only; all rights reserved.

Figure G1-18. (a) Uplift of precast concrete panels on pile-supported dock in Kao Lak Harbor, Thailand;
(b) uplift of precast floor slabs in coastal building in Patong Beach during the 2004 Indian Ocean tsunami.
Source: Saatcioglu et al. (2005).

Figure G1-19. (a) Three-span highway overpass in Rikuzentakata, Japan, with deck and approach
abutments washed away during the Tohoku tsunami (N 39.0087; E 141.6495); (b) multispan railway
bridge in Otsuchi, Japan, with girders washed away during the Tohoku tsunami (N 39.3587; E 141.9102).

Figure G1-20. Railway bridge piers in Otsuchi, Japan, that collapsed from hydrodynamic loads (N 9.3597;
E 141.91308).

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


12 Tsunami Loads and Effects

1.2  Tsunami Generation Mechanisms


Tsunamis can occur in any large body of water, including oceans, inland seas, and lakes.
Tsunamis can be generated by any mechanism that abruptly distorts the water surface.
The resulting restoration of equilibrium in the water surface manifests as tsunami waves
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 09/20/20. Copyright ASCE. For personal use only; all rights reserved.

radiating from the surface disturbance. The most common tsunami source mechanism is
the crustal plate uplift occurring during a subduction zone earthquake. Figure G1-21 illus-
trates how tectonic plate movement results in relative movement of the oceanic plate below
the continental plate at a subduction zone [Figure G1-21(a)]. Because of friction along the
interface between the plates, strain develops along the contact surface and the continental
plate is uplifted [Figure G1-21(b)]. When the accumulated strain exceeds the rock capacity,
the two surfaces slip relative to each other, causing an earthquake [Figure G1-21(c)]. This
release of strain results in sudden upward movement of the tip of the continental plate and

Figure G1-21. Subduction zone tsunami generation.


Source: Courtesy of USGS (2005).

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Introduction 13

simultaneous subsidence of the coastal zone [Fig-


ure G1-21(c)]. The upward movement of the sea floor Tsunami Generation
causes a deformation in the sea surface, which prop- Most of the recent trans-
agates away from the disturbance as tsunami waves oceanic damaging tsunamis have
resulted from subduction zone
[Figure G1-21(d)].
earthquakes, typically in the
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 09/20/20. Copyright ASCE. For personal use only; all rights reserved.

Because the subduction zone is generally ori- moment magnitude range of 8.0
ented parallel to the adjacent shoreline, tsunami and larger.
waves are focused toward the coastline and away
from the coastline. In deep ocean water, these tsu-
nami waves propagate at approximately 500 mph with extremely long wavelengths and
small amplitude (wave height). The wave energy propagates at different speeds depending
on the wavelength, resulting in dispersion of the tsunami energy into a series of waves that
approach the coastline. As the water depth decreases close to the shoreline, the wavelength
reduces and the amplitude increases, a process referred to as shoaling. The leading edge
of the tsunami may be either a positive inundation or a negative withdrawal, followed by a
series of inundation cycles depending on the extent of dispersion. If conditions are right,
such as a gradual sloping bathymetry (undersea topography) or shallow offshore reefs, the
tsunami waves may break before reaching the shoreline resulting in a series of tsunami bores
(Murata et al. 2010, Madsen et al. 2008). A tsunami bore is the turbulent leading edge of a
tsunami wave that results when the wave breaks. The leading edge of a tsunami wave may
experience fission resulting in a series of bores making up the full amplitude of the tsunami
wave. Alternatively, the tsunami may inundate the coastline as a surge or rapidly rising tide.
Resonance because of peculiarities in coastal bathymetry can extend the series of tsunami
waves for some time after the main tsunami energy has passed.
Most of the recent trans-oceanic damaging tsunamis have resulted from subduction zone
earthquakes, typically in the moment magnitude range of 8.0 and larger. This includes the Indian
Ocean tsunami of 2004, the Chile tsunamis of 2010 and 2015, and the Tohoku tsunami of 2011.
The second most common source mechanism for tsunamis is landslides occurring at
or below the water level. The rapid movement of a large volume of rock and soil material
during a landslide results in displacement of the surrounding water leading to tsunami waves
that radiate from the slide location. Because the amount of energy imparted to the water is
generally considerably less than during a subduction zone earthquake, landslide tsunamis
typically do not propagate very far from the source, although the wave amplitudes at the
source can be very large. The Lituya Bay tsunami in 1958 was generated by a landslide of an
estimated 30 million m3 entering the narrow inlet of Lituya Bay, Alaska. The landslide was
triggered by a 7.8-magnitude earthquake and caused a wave with amplitude estimated at 30
m and runup of 524 m on the other side of the inlet (Fritz et al. 2009). Once out of the bay,
however, the wave quickly dissipated.
Two recent tsunamis in Indonesia are attributed in whole or part to landslides. The Sunda
Strait tsunami on December 22, 2018, was most likely the result of a 150 to 200 million m3

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


14 Tsunami Loads and Effects

landslide resulting from the collapse of the southwest flank of Anak Krakatau, resulting in
more than 430 deaths on the shores of Java and Sumatra islands (Robertson et al. 2018; Grilli
et al. 2019). The effects of the tsunami were limited to the shores of Sunda Strait, where
Anak Krakatau is located. This volcano formed after the infamous 1883 eruption of Krakatoa
that resulted in a tsunami that caused more than 36,000 deaths (Wikipedia 2019a). The Palu
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 09/20/20. Copyright ASCE. For personal use only; all rights reserved.

tsunami on September 28, 2018, is attributed in part to a number of submarine landslides in


Palu Bay, Sulawesi, caused by a magnitude 7.5 earthquake (Robertson et al. 2019; Aránguiz et
al. 2020).
Subaerial and submarine landslides are also considered a potential contributor to the
tsunami hazard in fjords along the Alaskan shoreline (Suleimani et al. 2006) and the alluvial
deposits along the Southern California coastline (Bardet et al. 2003, Watts 2004).
Researchers are working on gaining a better understanding of the potential for earth-
quake-induced landslides to result in localized tsunamis along these and other coastlines.
Because the return interval and magnitude of these events are not well established, they
have not been included in the probabilistic tsunami hazard analysis (PTHA) performed as
part of the development of the ASCE 7-16 tsunami provisions.
Tsunamis can also result from earthquakes caused by strike-slip fault movement, but
typically only when the bathymetry is such that a large volume of water is displaced either
vertically or horizontally. In addition to the aforementioned landslide sources, the Palu,
Indonesia, tsunami is partly attributed to movement along the Palu-Koro Fault (Robertson
et al. 2019). The ASCE 7-16 tsunami design zone includes potential tsunamis generated by
strike-slip faults in Puget Sound because they present more of a danger to the Seattle-Ta-
coma region than the Cascadia Subduction Zone.
Trans-oceanic tsunamis attributable to large slope failures on the flanks of islands
(orders of magnitude greater than landslides) and asteroid strikes in the ocean have occurred
in geological time, but their recurrence intervals are too long and unpredictable for them to
be included in the ASCE 7-16 design standard.

1.3  Warning and Evacuation


Life safety during a tsunami cannot be guaranteed through coastal and structural engineer-
ing alone. It is essential that all coastal communities threatened by tsunamis develop, prac-
tice, and efficiently execute comprehensive tsunami warning and evacuation procedures so
as to reduce the loss of life as much as possible. The National Tsunami Warning Center in
Palmer, Alaska, and the Pacific Tsunami Warning Center in Honolulu, Hawai‘i, provide timely
information on all major earthquakes and any tsunamis that may result. These warnings are
relayed promptly to state agencies in all five western US states (Alaska, Washington, Oregon,
California, and Hawai‘i), as well as to other Pacific nations. It is then the responsibility of the
state and local emergency management agencies to coordinate message dissemination to
the public and coordinate evacuations, if required.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Introduction 15

It must be emphasized that tsunami design of


buildings and tsunami protective measures are not Tsunami Warning and
Evacuation
meant to replace the need for effective warning and
evacuation procedures for all tsunami-prone coast- Life safety during a tsunami
lines. The ASCE 7-16, Chapter 6 “Tsunami Loads and cannot be guaranteed
through coastal and structural
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 09/20/20. Copyright ASCE. For personal use only; all rights reserved.

Effects” provisions only apply to selected buildings


engineering alone. It is essential
in the inundation zone. The vast majority of resi- that all coastal communities
dential and other low-rise buildings will be severely threatened by tsunamis develop,
damaged, destroyed, or overtopped during a design- practice, and efficiently execute
comprehensive tsunami warning
level tsunami. It is therefore essential that the inun-
and evacuation procedures so as
dation zone be evacuated prior to arrival of the to reduce the loss of life as much
tsunami. Evacuation to high ground is always the as possible.
preferred approach. However, many communities
may not have adequate time in which to evacuate
all coastal residents to high ground prior to tsunami
Vertical Evacuation
inundation. In these communities, tsunami verti-
cal evacuation refuge structures designed to the By providing tsunami engineering
for all buildings that have a height
requirements of Chapter 6 will provide a high level
exceeding the tsunami inundation
of reliability against collapse during the maximum by a safe margin, a community
considered tsunami. In addition, taller reinforced can ensure that even those who
concrete or structural steel buildings that meet the do not evacuate according to
collapse prevention performance level according to plan can still survive a design-
level tsunami if they evacuate
Chapter 6 could also provide a refuge of last resort vertically.
for those caught in the tsunami zone without suffi-
cient time to reach high ground or a designated ver-
tical evacuation structure.
Having a long history of damaging tsunamis, Japan has developed an extensive warn-
ing and evacuation plan for all threatened coastal communities. Signage throughout the
tsunami inundation zone implores residents to evacuate to high ground or designated
evacuation structures (Figure G1-22). High-ground evacuation sites are most commonly
accessible by staircase, but in some instances ramps are also available (Figure G1-23).
Numerous buildings in the inundation zone are designated as tsunami vertical evacuation
refuges. Some of these are existing buildings evaluated as suitable for refuge (see Fig-
ures G1-9 and G1-24), whereas others were constructed specifically as tsunami refuges
(see Figure G1-7). During a tsunami warning, sirens and loudspeakers are used to keep the
population informed and encourage speedy evacuation. In spite of these valiant efforts,
more than 18,000 people died or are still missing after the Tohoku tsunami (National Police
Agency of Japan 2018, Wikipedia 2019b) in part because inundation was greater than antic-
ipated based on historical events, and defensive measures built at many coastal communi-
ties were inadequate to prevent inundation.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


16 Tsunami Loads and Effects
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 09/20/20. Copyright ASCE. For personal use only; all rights reserved.

Figure G1-22. (Left) Tsunami warning sign in Hachinohe, Japan (“Earthquake → Tsunami; Evacuation
→ “Run Away”); (right) map showing directions to evacuation buildings and high ground in Kamaishi,
Japan.

Figure G1-23. (Left) Tsunami evacuation ramp and stair access to high ground in Taro, Japan
(N 39.73685; E 141.97026); (right) stair access to high ground in Kamaishi, Japan (N 39.2753; E 141.8874).

In addition to those who evacuated to high ground or designated shelters, tens of thou-
sands of people survived the tsunami because they sought refuge in multistory buildings even
though they were not designated for use during a tsunami (Fraser et al. 2012). Although these
buildings and their occupants survived, many others lost their lives because they sought
refuge in buildings that were too low or could not survive the tsunami loads. By providing
tsunami engineering for all buildings that have a height exceeding the tsunami inundation
by a safe margin, a community can ensure that even those who do not evacuate accord-
ing to plan can still survive a design-level tsunami if they evacuate vertically. Communities
are therefore encouraged to require tsunami design for all Tsunami Risk Category (TRC) II
buildings that exceed a predetermined height, even though this is not strictly required by
ASCE 7-16, Chapter 6. The appropriate height for a particular community may vary from the

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Introduction 17
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 09/20/20. Copyright ASCE. For personal use only; all rights reserved.

Figure G1-24. Eight-story apartment building in Kamaishi, Japan, designated as a tsunami vertical evacuation
refuge with signage indicating safety at or above the fourth floor: (left) front and (right) rear.

65 ft recommended in ASCE 7-16, Commentary Section C6.1.1. This commentary section has


an extensive discussion on how a community might select the appropriate TRC II building
height above which tsunami design is required, and the ASCE Journal of Structural Engineer-
ing has a paper on design for tsunami resilience (Chock et al. 2018).

1.4  Development of ASCE 7-16, Chapter 6


The development of Chapter 6, “Tsunami Loads and Effects,” for inclusion in the 2016 version
of ASCE 7 officially began in February 2011 with the authorization of the Tsunami Loads and
Effects Subcommittee (TLESC) by the ASCE 7-16 Main Committee. However, the groundwork
for this effort was initiated much earlier. Figure G1-25 shows a timeline of the history of
tsunami design code development in the United States along with major damaging tsunami
events around the world.
The first and only jurisdiction in the United States to adopt tsunami design provisions
of any sort was the City and County of Honolulu on the island of Oahu, Hawai‘i (Honolulu
City & County 1984). This early tsunami code was based on a study by Dames & Moore (1980)
of prior laboratory research on tsunami loading. After a number of damaging tsunamis in
Japan and Southeast Asia, the National Tsunami Hazard Mitigation Program (NTHMP), in
conjunction with the National Oceanic and Atmospheric Administration (NOAA), organized
a workshop in November 2001 bringing together researchers “having expertise in structural,
marine, and civil engineering, seismology, geology, and emergency management to assess
the feasibility of (developing) construction guidelines and coastal land use guidance in areas

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


18 Tsunami Loads and Effects
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 09/20/20. Copyright ASCE. For personal use only; all rights reserved.

Figure G1-25. History of tsunami code development in the United States.

of both strong ground shaking and tsunami hazard” (Walsh et al. 2002). The NTHMP is a
coordinated US national effort to mitigate the impact of tsunamis through public education,
community response planning, hazard assessment, and warning coordination. The intent of
the workshop was to move tsunami research beyond the realm of modeling tsunami gener-
ation, propagation, and coastal inundation to include study of the effects of tsunamis on the
built environment, particularly after a major subduction zone earthquake as is anticipated
along the Cascadia Subduction Zone in North America. This workshop led to funding of a
research project to study the feasibility of designing coastal buildings to resist tsunami loads.
Harry Yeh at Oregon State University served as principal investigator (PI) on the project,
with Jane Preuss of Planwest Partners and the author as co-PIs. The result of this study was
published by Washington State Division of Geology and Earth Resources as Open File Report
2005-4 (Yeh et al. 2005). As part of this effort, Pacheco and Robertson (2005) investigated the
performance of three prototypical reinforced concrete building structures under three tsu-
nami conditions—namely, 3, 5, and 10 m flow depths at the building site. They concluded that
(1) “the prototype building with moment-resisting frame or dual system was able to resist the
tsunami forces”; (2) “the prototype building with shear wall-frame system was able to resist
the tsunami forces, however individual shear walls perpendicular to the tsunami flow may

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Introduction 19

fail and lead to progressive collapse of the building”;


and (3) bearing wall buildings, which utilize relatively “Tsunami Loads and Effects”
provisions were officially
thin walls to support all gravity and lateral loads, are
accepted as Chapter 6 in
likely to perform poorly during a tsunami event.
ASCE 7-16 on March 11, 2016,
The final report from this study was used as the exactly 5 years after the
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 09/20/20. Copyright ASCE. For personal use only; all rights reserved.

basis for an Applied Technology Council project (ATC Tohoku tsunami.


64) funded by FEMA and NOAA to develop guidelines
for design of structures for vertical evacuation from
tsunamis. The author participated in this effort as a member of the Project Management
Committee. The resulting guidelines were published in June 2008 as FEMA P-646, Guidelines
for Design of Structures for Vertical Evacuation from Tsunamis (FEMA 2008). These guidelines
were not written in mandatory language and relied on limited laboratory experiments to
infer tsunami flow conditions on land and the resulting structural loading.
Subsequent to review by the research community and the use of P-646 in preliminary
design of a vertical evacuation building in Cannon Beach, Oregon (Yu 2013), it was deter-
mined that portions of the guidelines needed to be updated, particularly relating to debris
impact loading. The author served as project technical director for the second edition of
FEMA P-646 published in April 2012 (FEMA 2012). In addition to providing guidance on tsu-
nami loading and design, FEMA P-646 also covers issues of siting, spacing, and sizing of
tsunami vertical evacuation structures. Although much of FEMA P-646 has been superseded
by ASCE  7-16, Chapter 6, there is still useful guidance to be gained from FEMA P-646 as
a community plans for designation of existing buildings or construction of new buildings,
structures, or mounds as vertical evacuation refuges. FEMA therefore funded a third edition
of FEMA P-646 that references ASCE 7-16 for most of the design aspects but retains useful
community planning guidelines (FEMA 2019). This third edition also incorporates community
planning guidance previously contained in FEMA P-646A, Vertical Evacuation from Tsuna-
mis: A Guide for Community Officials (FEMA 2009).
While working on the NTHMP research program, the author worked with colleagues
at the University of Hawai‘i at Manoa (UH Manoa), Oregon State University, and Princeton
University, and practicing structural engineers from Martin & Chock, Inc., in Honolulu to
develop a comprehensive research proposal titled “Development of Performance Based Tsu-
nami Engineering, PBTE” for submission to the National Science Foundation (NSF)-funded
George E. Brown Jr. Network for Earthquake Engineering Simulation (NEES). The project
was funded in 2005, shortly after the Indian Ocean tsunami on December 26, 2004 (Riggs
et al. 2005). This 5-year project included extensive tsunami experiments at the O. H. Hins-
dale Wave Research Laboratory at Oregon State University and the Hydraulics Laboratory at
UH Manoa. The results of this project were used to enhance tsunami inundation modeling,
scour, and sediment transport modeling and structural loading caused by tsunami bores.
After completion of this research project, Gary Chock of Martin & Chock, Inc., proposed the
formation of a subcommittee under the ASCE/SEI 7 Main Committee to transform these

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


20 Tsunami Loads and Effects

and other research results into tsunami design requirements for coastal construction. This
subcommittee consisted of 12 voting and 18 associate members with expertise in all areas of
tsunami engineering, including most of the co-PIs on the PBTE project.
The TLESC of ASCE 7-16 was officially approved by the ASCE Board of Directors in Feb-
ruary 2011 with Gary Chock as chair. The next month, on March 11, 2011, the Great East Japan
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 09/20/20. Copyright ASCE. For personal use only; all rights reserved.

Earthquake and subsequent Tohoku tsunami devastated many coastal communities on the
northeast coast of Honshu Island, Japan. One month after the tsunami, with funding from
ASCE, seven members of the ASCE 7-16 TLESC including the author, under the leadership of
Gary Chock, collaborated with tsunami researchers in Japan to perform an extensive survey
of structural and geotechnical damage caused by the tsunami (Chock et al. 2013b). The results
of this survey provided ideal opportunities to validate many of the laboratory experimental
results at full scale (Chock et al. 2013a). The lessons learned from this tragic event went a long
way to assisting the ASCE tsunami subcommittee in their development of tsunami design
provisions for ASCE 7-16.
Within three and a half years of its formation, the TLESC had developed a comprehen-
sive draft of tsunami design provisions for inclusion in ASCE 7-16. After another 18 months of
review, editing, and balloting by the ASCE 7-16 Main Committee and a 3-month public com-
ment period, the “Tsunami Loads and Effects” provisions were officially accepted as Chapter
6 in ASCE 7-16 on March 11, 2016, exactly 5 years after the Tohoku tsunami. ASCE 7-16 has
since been adopted by reference into the International Building Code (IBC) (ICC 2018), which
will form the basis for building codes in the four US states bordering the Pacific Ocean.
In November 2018, the ASCE 7-16, Chapter 6 “Tsunami Loads and Effects” provisions were
adopted by amendment into the Hawai‘i State Building Code, making Hawai‘i the first US
state to officially adopt the new tsunami design provisions.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


2
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

Prototype Buildings

Two reinforced concrete prototypical building structures were developed for use in the design
examples demonstrated in this guide. The initial layout and structural configurations of these
buildings were determined with the assistance of Gary Chock of Martin & Chock, Inc., a Hono-
lulu structural engineering consulting company, to ensure that they represent realistic midrise
office and residential buildings. These Tsunami Risk Category II buildings were developed with
sufficient height above grade (more than 65 ft as suggested for severe tsunami hazard regions
in the commentary to ASCE 7-16, Chapter 6) to provide a refuge of last resort for those caught
in the tsunami inundation zone without sufficient time to evacuate to high ground, and to
increase community resilience by ensuring that larger buildings remain intact except for non-
structural components at the lower levels. They were located near the shoreline and with their
broad dimension facing the incoming tsunami flow. Figures G2-1 and G2-2 show the Seaside,
Oregon, and Waikiki, Hawai‘i, locations in images from the ASCE 7 Tsunami Design Geodata-
base, respectively. Two other sites at Monterey, California, and Hilo, Hawai‘i, were also consid-
ered (McKamey and Robertson 2019). The Seaside location was selected to represent a coastal
community in the US Northwest exposed to the Cascadia Subduction Zone. The Monterey
location was selected as a typical California coastal community that is not directly exposed to
the Cascadia Subduction Zone, but rather to distant-source tsunamis. The Hilo location was
selected because of the extremely high flow depths anticipated at that location. The Waikiki
location on O’ahu was selected because it is one of the few locations in the five western US
states that does not require high seismic design. As a result, the buildings are not as strong or
ductile as equivalent buildings in the other three locations.
These prototype buildings are used in examples throughout this guide. In addition, this
guide’s appendix provides a detailed design example for the Seaside location. Full building
design examples for both prototype buildings at all four locations are presented by McKamey
and Robertson (2019).
Structural steel buildings are also expected to perform well during tsunami loading if
designed to the provisions of ASCE 7-16. Although both examples illustrated in this guide are
reinforced concrete buildings, this does not imply that structural steel buildings cannot be
designed for tsunami loads and effects.

21
@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


22 Tsunami Loads and Effects
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

Figure G2-1. Seaside location for prototype building.


Source: ASCE Tsunami Design Geodatabase (2016).

Figure G2-2. Waikiki location for prototype building.


Source: ASCE Tsunami Design Geodatabase (2016).

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Prototype Buildings 23

2.1  Reinforced Concrete Moment-Resisting Frame


Office Building
A 6-story office building consisting of reinforced concrete moment-resisting frames (MRFs),
a flat plate post-tensioned concrete floor system, and interior gravity load columns was
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

analyzed and designed for wind and seismic loading conditions specified by ASCE 7-16 for
each site. Figure G2-3 shows the floor plan and section for this building. This prototype
building was analyzed and designed for the ASCE 7-10 specified wind and seismic condi-
tions at two locations: Waikiki, Hawai‘i; and Seaside, California (Yokoyama and Robertson
2014, McKamey and Robertson 2019). The Seaside building has perimeter and interior special
moment-resisting frames, whereas the Waikiki building has perimeter and interior interme-
diate moment-resisting frames.

Figure G2-3. Prototype 6-story reinforced concrete MRF office building: Plan and section views.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


24 Tsunami Loads and Effects

2.2  Reinforced Concrete Shear Wall


Residential Building
The second prototype building is a 7-story residential building consisting of reinforced con-
crete shear walls (SWs) at elevator shafts and stairwells, a flat plate post-tensioned concrete
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

floor system, and reinforced concrete gravity load columns. Figure G2-4 shows the floor plan
and section for the residential building. This prototype building was analyzed and designed
for the wind and seismic conditions specified by ASCE 7-10 at two locations: Waikiki, Hawai‘i;
and Seaside, Oregon (Yokoyama and Robertson 2014, McKamey and Robertson 2019). The
residential building in Seaside has special reinforced concrete shear walls, and the Waikiki
building has ordinary reinforced concrete shear walls. It would be more common for the
elevator shafts and stairwells to be located in the interior of the building, but here they are
moved to the exterior so as to demonstrate the tsunami loading effects on exterior struc-
tural wall elements.

Figure G2-4. Prototype 7-story reinforced concrete shear wall residential building: Plan and section
views.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Prototype Buildings 25

2.3  Seismic and Wind Design


The prototype buildings were located in four coastal communities with varying wind, seis-
mic, and tsunami loading conditions. Table G2-1 lists the locations with associated longitude,
latitude, and design wind speeds according to ASCE 7-10. Table G2-2 lists the seismic design
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

criteria at each site assuming Soil Classification D for “stiff soil.” The Waikiki site has a higher
design wind speed but lower seismic demand than the other sites.
Each prototype building was analyzed and designed for ASCE 7-10 wind and seismic
conditions (Yokoyama and Robertson 2014). They were then evaluated for the tsunami loads
required by ASCE 7-16 for these locations (McKamey and Robertson 2019).

Table G2-1. Prototype Building Locations and Associated Design Wind Speeds.

Location Latitude Longitude Design wind speed (mph)

Monterey, CA N 36.6002 W 121.8818 110

Seaside, OR N 45.9948 W 123.9295 110

Hilo, HI N 19.7209 W 155.0833 130

Waikiki, HI N 21.2755 W 157.8255 130

Table G2-2. Prototype Building Seismic Design Criteria.

Seismic design criteria

Seismic
design
Location Site class category SS S1 SDS SD1

Monterey, CA D D 1.513g 0.554g 1.009g 0.554g

Seaside, OR D D 1.332g 0.683g 0.888g 0.683g

Hilo, HI D D 1.500g 0.600g 1.000g 0.600g

Waikiki, HI D C 0.579g 0.170g 0.516g 0.240g

Table G2-3 lists the seismic design parameters for all four sites considering Site Soil
Classification D. According to ASCE 7-10, all the buildings fall into Seismic Design Cate-
gory D. However, the Honolulu Building Code amendments modify ASCE 7-10, Tables 11.6-1
and 11.6-2 so that Risk Category I and Risk Category II buildings can be designed as Seismic
Design Category C if SDS<0.6g and SD1<0.25g (Honolulu City & County 2012). Because of this
local amendment, the Waikiki location represents a lower level of seismic design and detail-
ing than the other locations.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


26 Tsunami Loads and Effects

Each prototypical building was designed for dead, live, wind, and seismic loads follow-
ing ASCE 7-10 (ASCE 2010) standard design requirements. The moment-resisting systems
were designed using the equivalent lateral force approach and optimized to just satisfy the
interstory drift limits required for the seismic design. Each building was then designed and
detailed following ACI 318-14 (2014) seismic design provisions. These prototype buildings will
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

be used in the tsunami design examples in this guide and as a full worked example of the
Seaside location in the appendix.

Table G2-3. Reinforced Concrete Prototype Building Seismic Design Parameters.

Prototype building Seismic design base shear


location Eh (kips)

Office building with Residential building with


moment-resisting frames shear walls

Seismic Seismic Seismic


design weight Eh weight Eh
Location category (kips) Ta (s) Ra Ωo Cs (kips) (kips) Ta (s) R
b
Ω0 Cs (kips)

Monterey, CA D 25,075 0.77 8 3 0.0899 2,254 19,633 0.463 6 2.5 0.168 3,298

Seaside, OR D 25,075 0.77 8 3 0.1109 2,781 19,633 0.463 6 2.5 0.148 2,906

Hilo, HI D 25,075 0.77 8 3 0.097 2,432 19,633 0.463 6 2.5 0.167 3,279

Waikiki, HI C 25,075 0.77 5 3 0.0623 1,562 19,633 0.463 5 2.5 0.1032 2,026
a
Special moment frames in Monterey, Seaside, and Hilo; intermediate moment frames in Waikiki.
b
Special shear walls in Monterey, Seaside, and Hilo; ordinary shear walls in Waikiki.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


3
General Requirements
Downloaded from ascelibrary.org by La Trobe University on 09/21/20. Copyright ASCE. For personal use only; all rights reserved.

The tsunami design provisions in Chapter 6 of ASCE 7-16 are arranged in a sequence such
that they can be applied sequentially by the design engineer. The flowcharts in Figures G3-1
and G3-2 show an outline of this process. The section numbers in the flowcharts refer to the
relevant sections in ASCE 7-16, but also correspond to the chapters in this guide. For exam-
ple, ASCE 7-16, Section 6.4 on tsunami risk categories will be covered in Chapter 4 of this
guide, and Section 6.5 will be covered in Chapter 5 of this guide, and so forth.
Tsunami design is only required for buildings and structures falling into one of the TRCs
specified by the local jurisdiction. ASCE 7-16 requires tsunami design for TRC III and IV build-
ings, and its commentary recommends that the local jurisdiction consider requiring tsunami
design for TRC II buildings of sufficient height to serve as refuges of last resort for those
stranded in the inundation zone and for increased disaster resilience of the community.
Greater detail on the tsunami risk categories is given in Chapter 4 of this guide.
Tsunami design is required if the building or structure falls within the mapped tsunami
design zone. An energy grade line analysis (EGLA) must be performed to determine the max-
imum flow depth and maximum flow velocity at the project site. This process is described in
Chapter 6 of this guide. For TRC IV buildings, a site-specific PTHA must also be performed
and the results of this analysis must be compared with the EGLA results to avoid possible
modeling underestimation of the flow velocity. A site-specific PTHA can also be performed
for TRC II and III buildings but is not required. The requirements for a site-specific PTHA are
outlined in Chapter 7 of this guide.
Chapter 8 covers the structural design procedures outlined in ASCE 7-16, Section 6.8,
including load cases, load combinations, and structural acceptance criteria. Chapters 9, 10,
and 11 cover hydrostatic, hydrodynamic, and debris impact loading, respectively, for which
the structural system and components must be designed. Chapter 12 covers foundation
design for unique geotechnical aspects of tsunami design, including general erosion, scour,
and seepage effects. Countermeasures to protect foundations from tsunami effects are also
addressed. Chapter 13 provides information on acceptable tsunami countermeasures includ-
ing tsunami barriers. Chapter 14 addresses additional requirements for the design of build-
ings and structures that are specifically designated as tsunami vertical evacuation refuge

27
@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


28 Tsunami Loads and Effects
Downloaded from ascelibrary.org by La Trobe University on 09/21/20. Copyright ASCE. For personal use only; all rights reserved.

Figure G3-1. Flowchart showing ASCE 7-16 tsunami design sequence: Part A.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


General Requirements 29
Downloaded from ascelibrary.org by La Trobe University on 09/21/20. Copyright ASCE. For personal use only; all rights reserved.

Figure G3-2. Flowchart showing ASCE 7-16 tsunami design sequence: Part B.

structures. Because of the potential for loss of life if these structures fail or are overtopped
during a tsunami, the requirements in this section significantly enhance the reliability of the
structural performance during a maximum considered tsunami. Finally, Chapter 15 addresses
the tsunami requirements for designated nonstructural components and systems, and Chap-
ter 16 covers tsunami design of non-building structures.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


30 Tsunami Loads and Effects

3.1  Scope
3.1.1  Geographic Extent of Applicability
The 2016 edition of the ASCE 7 “Tsunami Loads and Effects” chapter is applicable only to
the five US states bordering the Pacific Ocean—Alaska, Washington, Oregon, California, and
Hawai‘i. Coastal regions in these five states are exposed to a quantifiable probabilistic tsu-
nami hazard resulting from subduction zone earthquakes. The MCT is defined as having a
Downloaded from ascelibrary.org by La Trobe University on 09/21/20. Copyright ASCE. For personal use only; all rights reserved.

2% probability of being exceeded in a 50-year period. This represents a 1:2,475 annual proba-
bility of exceedance, often referred to as the 2,500-year mean return interval (MRI) tsunami.
The 2,500-year MRI was selected to be consistent with the ASCE seismic design provisions,
which are based on a maximum considered earthquake (MCE) with the same 2,500-year MRI.
In ASCE 7-16, Chapter 6, the 2,500-year MCT is represented by the offshore tsunami
amplitude and characteristic period at 100 m bathymetric depth with respect to mean sea
level along the entire coastline of Washington, Oregon, California, and Hawai‘i, and most of
the southern coastline of Alaska. These data are available for download from the ASCE Tsu-
nami Hazard Tool (http://asce7tsunami.online). Figure G3-3 shows an example of the MCT

Figure G3-3. Example of the offshore wave amplitude and period data points along the 100 m
bathymetric line offshore from Seaside, Oregon. Each of the mapped points has the wave amplitude,
period, and exact location for that data point, as well as a map showing the disaggregated sources
contributing to the 2,500-year probabilistic tsunami at that location.
Source: ASCE Tsunami Design Geodatabase (2016).

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


General Requirements 31

wave data offshore from Seaside, Oregon. The extent of coverage of the ASCE 7-16 offshore
tsunami maps is indicated in ASCE 7-16, Figure 6.1-1. A portion of the southern coastline of
Alaska was not included because of the lack of adequate resolution bathymetry and topog-
raphy. In addition, the northern and western Alaska coastlines were not included because
they are not directly exposed to tsunamis generated by subduction zone earthquakes and
presently lack adequate data to perform a probabilistic tsunami hazard analysis. For similar
reasons of data availability, other US territories such as Puerto Rico, US Virgin Islands, Guam,
Northern Mariana Islands, and American Samoa were also not included in this first version
Downloaded from ascelibrary.org by La Trobe University on 09/21/20. Copyright ASCE. For personal use only; all rights reserved.

of the MCT mapping. The commentary to Chapter 6 recommends that for Tsunami Risk
Category III and IV buildings in these locations, consideration be given to performing prob-
abilistic site-specific tsunami hazard analysis for use in implementing the tsunami design
provisions.
As mentioned in Chapter 1, subduction zone earthquakes are not the only possi-
ble source for generating tsunamis. The ASCE 7-16 tsunami design zone maps include the
effects of three potential local faults in Puget Sound—the Seattle Fault, Tacoma Fault, and
Rosedale Fault (Wei et al. 2017). Local tsunamis can also be generated by large landslides at
or near the coast and by underwater volcanic eruptions. In particular, Alaska has a history
of co-seismic landslides generating local tsunamis in fjords. Because of a lack of information
on likely return periods for these landslide-induced tsunamis, they were not included in the
ASCE 7-16 Tsunami Design Zone maps.

3.1.2  Applicability to Buildings and


Other Structures
The ASCE 7-16 “Tsunami Loads and Effects” chapter requires that tsunami design be applied
to all TRC IV structures located within the tsunami design zone (TDZ) and to all TRC III struc-
tures with inundation depth exceeding 3 ft at any point within the building footprint. In the
commentary, the local jurisdiction is encouraged to require tsunami design for TRC II build-
ings with sufficient height to provide emergency refuge for people stranded within the TDZ.
This will also enhance the resilience of the community since these buildings are likely to be
structurally sound after the tsunami, requiring only replacement of nonstructural elements
below the maximum tsunami flow elevation for them to resume their pre-tsunami function.
Short TRC II buildings such as light-framed single- and multiple-family residences and
low-rise office buildings and those that are only a single story (like warehouses and airplane
hangars) are not required to be designed for tsunami loading. In general, the economic cost
of hardening these more vulnerable buildings for tsunami loads is not justified considering
the inability of these low-rise or single-story structures to provide safe refuge for people
stranded in the tsunami zone or any postinundation functionality.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


32 Tsunami Loads and Effects

3.2  Probabilistic Tsunami Hazard Analysis


All the Pacific basin tsunami sources were considered in development of the 2,500-year MRI
probabilistic offshore tsunami wave conditions for the ASCE 7-16 design standard. Past reli-
ance on historical tsunamis to define the design level has failed, most notably during the
Tohoku tsunami in Japan, so a probabilistic approach is the preferred methodology to more
reliably estimate the offshore wave conditions of future potential tsunamis to achieve target
reliabilities against failure.
Downloaded from ascelibrary.org by La Trobe University on 09/21/20. Copyright ASCE. For personal use only; all rights reserved.

The potential tsunami sources were fully inte-


grated over seismic sources varying in size and recur-
Probabilistic Tsunami
rence rate. Uncertainties were included through the
Hazard Analysis, PTHA
use of logic trees and probability distribution func-
tions based on earthquake source predictions by the Past reliance on historical
tsunamis to define the design
US Geological Survey (USGS). The results include
level tsunami has failed, so a
aleatory uncertainties relating to the tsunami mod- probabilistic approach is the
eling accuracy, particularly for nonlinear coastal preferred methodology to more
inundation, and epistemic uncertainties in the tec- reliably estimate the offshore
tonic fault maximum magnitude, slip distribution, wave conditions of future
potential tsunamis.
recurrence, and coupling ratio parameters (Thio et
al. 2017). The resulting probabilistic tsunami hazard
is primarily expressed in the offshore tsunami ampli-
tude and characteristic period at 100 m bathymetric depth relative to mean sea level along
the entire coastline of Washington, Oregon, California, and Hawai‘i, and the southern coast-
line of Alaska. Although tsunamis are caused by subduction zone earthquakes, the hazard
results from tsunami wave runup causing on-land flooding. Therefore, aleatory uncertainty
in the tsunami inundation modeling is accounted for by applying a lognormal distribution
function to these offshore results based on the statistical difference of measured runups to
actual runups for past events. During generation of the hazard-consistent 2,500-year off-
shore wave properties, source disaggregation was recorded for use in generating scenario
tsunamis that represent the 2,500-year MRI event. An example of the disaggregated tsunami
hazard for Waikiki, Hawai‘i, is shown in Figure G3-4. These data are stored in the online
ASCE Tsunami Design Geodatabase. The probabilistic tsunami hazard analysis conducted to
support ASCE 7-16 is described in more detail in Thio et. al. (2017).

3.3  Tsunami Design Zone Maps


The final step of the PTHA process involves the inundation analysis to determine the tsunami
runup on land. The runup is the ground elevation where the tsunami inundation reaches its
horizontal limit. The ASCE Tsunami Design Geodatabase includes these geocoded points
defining the locations and elevations of the runup. The inundation limit on land is the

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


General Requirements 33
Downloaded from ascelibrary.org by La Trobe University on 09/21/20. Copyright ASCE. For personal use only; all rights reserved.

Figure G3-4. Disaggregated tsunami hazard for Waikiki, Hawai‘i.

smoothed extent line formed by these discrete runup points, and the tsunami design zone
essentially consists of the land area between the inundation limit line and the coastline.
To determine the extent of tsunami inundation along the exposed coastlines, the off-
shore wave heights were used to generate onshore inundation modeling to determine the
TDZ and runup elevation. The disaggregated sources were used to generate hazard-consis-
tent scenario tsunamis, which were propagated toward the coastline of interest. The result-
ing offshore wave amplitudes generated by the hazard-consistent scenario tsunamis at the
100 m bathymetric depth matched, on average, the ASCE 2,500-year wave amplitudes, but
no point was less than 80% of the ASCE 2,500-year offshore wave amplitudes anywhere
along the modeled coastline.
The tsunami was then propagated on land using a 30 or 60 m horizontal grid resolu-
tion digital elevation model (DEM) to determine the inundation limit. All land seaward of the
inundation limit is in the TDZ. This work was performed by researchers at the NOAA Pacific
Marine Environmental Laboratory (PMEL) in Seattle, Washington, and is described in more
detail in Wei et. al. (2017) (Review G3-1).
Although the 60 m grid is relatively coarse, it was adequate to develop the TDZ for the
entire US Pacific Ocean coastline being mapped. It is recommended that the five individual
US states use higher-resolution DEMs to generate more accurate TDZ inundation limits for
highly populated areas. (These data sets may then be incorporated into future editions of the
ASCE 7 tsunami design provisions.) At the time of writing this guide, both Hawai‘i and Cali-
fornia were in the process of re-creating refined extents of the tsunami design zones using
10 m horizontal resolution DEMs (Review G3-2).

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


34 Tsunami Loads and Effects

Review G3-1. Tsunami Model Terminology.

Tsunami modeling is generally performed using a two-dimensional (2D) hydraulic


model. This model uses a full three-dimensional (3D) digital representation of the
bathymetry (topography of the ocean floor) and topography (DEM) so that flow is free
to propagate in any horizontal direction.
The most common numerical models assume that the water has no vertical veloc-
Downloaded from ascelibrary.org by La Trobe University on 09/21/20. Copyright ASCE. For personal use only; all rights reserved.

ity, so there are only two horizontal velocity components. In addition, it is assumed
that any variation in the horizontal velocity through the flow depth can be repre-
sented by a single velocity by integrating over the flow depth. This is referred to as
a “depth-averaged 2D” numerical model, or colloquially as a hydrostatic model. Nei-
ther of these assumptions is absolutely correct, but they are acceptable for deep open
ocean flow and, to a lesser extent, for coastal and overland flow. Clearly rapid changes
in the bathymetry such as presented by a steep continental shelf or a fringing reef will
result in vertical velocities in the flow. In addition, friction between the water and the
ground surface will result in a flow velocity profile through the water depth. These
phenomena are not captured precisely by a hydrostatic model. In addition, hydrostatic
models do not generate wave dispersion, so trans-oceanic tsunami waves are domi-
nated by a single large wave as opposed to the more realistic tsunami wave train.
Higher-order numerical models are available that include the vertical velocity,
leading to wave dispersion for trans-oceanic tsunamis. Tsunami models based on Bous-
sinesq equations also generate dispersion of the various wave frequencies contained in
the tsunami, resulting in realistic tsunami wave trains. Additional higher-order effects
such as consideration of gravitational variability, water compressibility, earth elastic-
ity, and so forth have less effect on the waveform but can correct for slight errors in
tsunami wave arrival times. Typically, these higher-order models require considerably
greater computational time and are therefore less commonly used by tsunami mod-
elers. Fully 3D hydraulic models (similar to 3D solid models) do exist but are typically
only used in research for small domains around a structure of interest.
The major advantage of using hydrostatic 2D tsunami numerical models is the
significant reduction in computational time required to run the tsunami analysis.

3.4  Definitions, Symbols, and Notation


Because design for tsunami loading may be new to many users of this guide, it is strongly recom-
mended that the reader review all the definitions provided in ASCE 7-16, Section 6.2 and the sym-
bols and notation presented in Section 6.3. Every attempt was made by the Tsunami Loads and
Effects Subcommittee to utilize tsunami terminology and symbols that are consistent with the
latest tsunami glossary published by the Intergovernmental Oceanographic Commission (2013).

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


General Requirements 35

Review G3-2. Digital Elevation Model Resolution.

The resolution of a DEM refers to the horizontal spacing of elevation data points in
the DEM. A resolution of 1/3 s implies that the ground elevation is reported at every
1/3 arcsecond (arc-s) of latitude and longitude.
With the circumference of the earth approximately 40,000,000 m, a 1 arc-s reso-
lution represents a distance of
Downloaded from ascelibrary.org by La Trobe University on 09/21/20. Copyright ASCE. For personal use only; all rights reserved.

S = 40,000,000/(360° × 60 min × 60 s) = 30.9 m

A resolution of 1/3 arc-s therefore represents elevation data points at approxi-


mately 10 m spacing in each direction at the equator. Away from the equator the spac-
ing in the east-west direction reduces as the lines of longitude converge.
In general, the vertical resolution of the DEMs provided in the ASCE Tsunami
Design Geodatabase is on the order of a few inches.

Figure G3-5 is a copy of ASCE 7-16, Figure 6.2-1


illustrating some of the key definitions relating to appli- Tsunami terms and
definitions may not be
cation of the tsunami design provisions. In particular,
familiar to many users.
it is important to understand the terms inundation
limit and runup. Inundation limit is the horizontal dis- It is important to review all
definitions, symbols, and
tance from the geoid reference shoreline to the maxi-
notation relating to Chapter 6
mum inland extent of the tsunami inundation, whereas so as not to misinterpret the
runup, R, is the elevation of the maximum inundation tsunami design provisions.
point above the sea level datum. The water level at the

Figure G3-5. Some of the key definitions used throughout the tsunami design provisions.
Source: ASCE 7-16, Figure 6.2-1.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


36 Tsunami Loads and Effects

site of interest can be referenced in one of two ways. One can refer to either the “inundation
elevation,” which refers to the water level above the geoid reference sea level elevation, or
the “inundation depth,” which refers to the depth of water above grade at the site. This latter
definition is the flow depth used for loading and scour calculations. The inundation depth is
equal to the inundation elevation minus the ground elevation at the site.
Downloaded from ascelibrary.org by La Trobe University on 09/21/20. Copyright ASCE. For personal use only; all rights reserved.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


4
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

Tsunami Risk Categories


ASCE 7-16 specifies Risk Categories (RC) I through IV for all buildings and other structures. In
ASCE 7-16, Chapter 6, some modifications are made to these standard risk categories resulting in
Tsunami Risk Categories I through IV.
RC IV buildings are all essential facilities that
should survive the design-level loads without signifi- Tsunami risk categories
cant damage so that they can easily recover and con- are different from Standard
tinue their pre-event function. Typical RC IV buildings ASCE 7 risk categories!
include hospitals, emergency management centers, It is important to review
fire stations, and ambulance facilities, among others. the local code amendments
For the purposes of tsunami design, it is important applicable to a particular
that these essential facilities survive the event and jurisdiction to establish the
appropriate tsunami risk
are available to the community immediately thereaf- category for design of each
ter. However, some RC IV facilities will be evacuated building or structure.
during the tsunami, such as police and fire stations,
ambulance facilities, and emergency aircraft hangars.
It is not essential that these buildings survive the event, as long as the emergency personnel
and their equipment have evacuated to a safe location and are available to serve the commu-
nity immediately after the event. Requiring tsunami design of these facilities would create
an unnecessary economic hardship that would tend to result in relocation of the facilities
outside of the tsunami design zone. This would disadvantage communities in the TDZ during
all other events requiring police, fire, and ambulance services. They are therefore removed
from TRC IV and assigned to either TRC II or III at the discretion of the local jurisdiction.
Other essential facilities such as earthquake and hurricane shelters, which may be located
in the TDZ, will be evacuated during the tsunami warning, and they will not be expected to
survive undamaged. Alternate shelters outside the TDZ will need to be designated for refuge
during the tsunami and post-tsunami sheltering of those unable to return to their homes in
the TDZ.
TRC III buildings are generally high-occupancy facilities such as auditoriums, schools,
and other major educational centers, as well as custodial facilities such as jails and detention

37
@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


38 Tsunami Loads and Effects

centers. In addition, the local jurisdiction may elect to include critical facilities such as pow-
er-generating stations, water treatment facilities for potable water, wastewater treatment
plants, and other public utility facilities not included in TRC IV. This will greatly enhance the
resilience of the community and accelerate post-tsunami recovery (Review G4-1).
TRC I includes buildings and other structures that are generally unoccupied or are
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

unlikely to result in loss of life or injury if they were to collapse during a design-level event.
TRC II buildings include everything that is not designated as TRC IV, III, or I. This is the
vast majority of buildings such as office, residential, hotel, and other mid- to high-rise build-
ings, as well as all single- and multifamily residences. As discussed in Section 3.1, “Scope,”
most Risk Category II buildings are exempt from the ASCE 7-16 tsunami design requirements
except for those of sufficient height to make tsunami design practically achievable and to
serve as a refuge of last resort. The appropriate height will be designated by the local juris-
diction, so it is important to check local amendments applicable to the project site.
ASCE 7-16 leaves the decision on the appropriate TRC for certain categories of building
up to the local jurisdiction during the code adoption process. It is therefore important to
review the local amendments applicable to a particular jurisdiction to establish the appro-
priate TRC for each building or structure design.

Review G4-1. Critical Facilities.

During the Tohoku tsunami, the Minamigamou Wastewater Treatment Plant in Miyagi
Prefecture, which provided primary and secondary treatment for 70% of the waste-
water from 1 million people living in Sendai, Japan, was extensively damaged. For one
month after the tsunami, all wastewater from Sendai was discharged directly to the
ocean with no treatment whatsoever. It took a month to reestablish a basic prima-
ry treatment operation, which was the only treatment provided during the 4 years
required to completely rebuild the wastewater treatment plant. Although the plant
was rebuilt at the same location, all critical equipment has now been elevated above
the March 11, 2011, tsunami inundation elevation, or protected from water damage if at
a lower elevation.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


5
Tsunami Flow Characteristics
Downloaded from ascelibrary.org by La Trobe University on 09/21/20. Copyright ASCE. For personal use only; all rights reserved.

Tsunami flow characteristics such as inundation depth and flow velocity are essential for
all aspects of tsunami design. This chapter provides the background to development of the
methods provided in ASCE 7-16 for determining flow depth and velocity at the project site.

5.1  TRC II and III Buildings and Other Structures


The primary procedure for determining flow depth and velocity is the Energy Grade Line
Analysis (EGLA) developed by members of the ASCE  7-16 TLESC specifically for use in
ASCE 7-16 (Kriebel et al. 2017). This method uses the runup elevation at the inundation limit
shown in the TDZ maps to develop a profile of the maximum flow depth and velocity between
that point and the coastline. Chapter 6 provides a detailed explanation of EGLA with suitable
examples so that the reader can understand how to generate their own spreadsheet or soft-
ware code to apply the EGLA to any location within the mapped tsunami design zone. EGLA
is considered adequate for all TRC II and III buildings and other structures located in the TDZ
that require tsunami design. Both EGLA and site-specific tsunami inundation modeling are
typically required for TRC IV structures as discussed in Section 5.2.

5.1.1  Runup Evaluation for Areas Where No Map


Values Are Given
Tsunami design zone maps have been generated following the procedure described in Sec-
tion 6.7 for all coastlines of Washington, Oregon, California, and Hawai‘i, and for most of the
southern coastline of Alaska. The full extent of TDZ coverage is shown in ASCE  7-16, Fig-
ure 6.1-1. When TDZ maps are not available for a particular location, then EGLA requires an
alternate means to determine the runup elevation and inundation limit. A simple procedure
is provided in ASCE 7-16, Section 6.5.1.1 to conservatively estimate runup elevation using the
surf similarity parameter based on the offshore wave height given in ASCE 7-16, Figure 6.7-
1. The background to this procedure is explained in detail in the ASCE  7-16 Commentary.
Examples G5-1 and G5-2 show the application to typical coastal sites. This approach might

39
@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


40 Tsunami Loads and Effects

be useful for US territories threatened by tsunamis that are not included in the ASCE 7-16
TDZ maps, and for other countries that have not yet developed PTHA inundation maps.
Examples G5-1 and G5-2 show that this approach provides results with considerable con-
servatism. The surf similarity approach represents an upper bound of both laboratory and
field data as shown in ASCE 7-16, Figure C6.5-3.
Downloaded from ascelibrary.org by La Trobe University on 09/21/20. Copyright ASCE. For personal use only; all rights reserved.

Example G5-1. Surf Similarity Example 1.

Surf similarity is a parameter commonly used in coastal engineering.


The surf similarity parameter based on offshore wave parameters at the 100 m
bathymetric contour is given as

TTSU g
100  (6.5-1)
cot 2 H T

where φ is the mean slope angle of the nearshore profile taken from the 100 m water
depth to coastline defined by the mean high-water elevation along the axis of the
topographic transect for the site, and TTSU and HT are the offshore wave period and
amplitude, respectively, and are given in the ASCE Tsunami Design Geodatabase
(https://asce7tsunami.online/). The period, TTSU, is given in minutes in the ASCE Tsu-
nami Design Geodatabase, so must be converted to seconds before substituting into
Equation (6.5-1).
For the Seaside, Oregon, site (N 45.99; W 123.93), the distance from the 100 m
bathymetric line to the shoreline can be determined using the measurement tool in
the ASCE Tsunami Design Geodatabase as 25,260 m, resulting in a mean slope angle of
 100 
f = arctan   = 0.23o
 25, 260 
Therefore, cot(ϕ) = 253 = 25, 260/100, as expected. From the ASCE Tsunami Design
Geodatabase, HT = 28 ft and TTSU = 20 min = 1,200 s.

1, 200 32.2
Therefore, 100   2.03  0.6 but < 6
253 2  28

 R
H T  2.50 log10 100    2.05  2.50 log10  2.03   2.05  2.82

from ASCE 7-16, Equation (6.5-2b) and Figure 6.5-1.


So, R = 2.82HT = 2.82 × 28 = 79 ft, compared with 55.55 ft from the ASCE Tsunami
Design Geodatabase.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Tsunami Flow Characteristics 41

Example G5-2. Surf Similarity Example 2.

For the Waikiki, Hawai‘i, site (N 21.276; W 157.826), the distance from the 100 m bathy-
metric line to the shoreline can be determined using the measurement tool in the
ASCE Tsunami Design Geodatabase as 2,575 m, resulting in a mean slope angle of
 100 
f = arctan   = 2.22o
 2, 575 
Downloaded from ascelibrary.org by La Trobe University on 09/21/20. Copyright ASCE. For personal use only; all rights reserved.

2, 575
Therefore, cot(f) = 25.75 = , as expected.
100
From the ASCE Tsunami Design Geodatabase, HT = 18 ft and TTSU = 15 min = 900 s.

900 32.2
Therefore, 100   18.65 > 6 but < 20
25.75 2 18

R
H T  4.0 from ASCE 7-16, Equation (6.5-2a).
So, R = 4HT = 4 × 18 = 72 ft, compared with 23 ft from the ASCE Tsunami Design
Geodatabase.

5.2  TRC IV Buildings and Other Structures


For Tsunami Risk Category IV buildings and other structures, the EGLA described in Chap-
ter 6 is performed to determine the maximum flow depth and velocity at the project site. If
the flow depth determined using EGLA exceeds 12 ft, then a site-specific probabilistic tsu-
nami hazard analysis must also be performed. The procedure for performing this analysis
is described in Chapter 7. A site-specific PTHA will yield complete time histories for flow
depth (h), flow velocity (u) vectors, and momentum flux (hu2). Although site-specific PTHA is
only required for TRC IV buildings and other structures, it can be used for design of any TRC
building and other structures, if desired (Figure G5-1).
The flow depth determined by site-specific PTHA can be used directly for the tsunami
design. However, the flow velocity determined by site-specific PTHA must satisfy the limits
specified in ASCE 7-16, Section 6.7.6.8 when compared with the results of an energy grade
line analysis for the same location. ASCE  7-16, Section 6.7.6.8 requires that the maximum
velocity determined by site-specific PTHA not be less than 90% of the velocity determined
by EGLA for urban environments (using Manning’s coefficient of 0.04) and not less than 75%
of the velocity determined by EGLA for other terrain roughness. This protects against mod-
eling errors or underestimation of the uncertainty in the site-specific analysis and ensures
conformity of the design to the ASCE  7-16 provisions, thereby maintaining the desired

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


42 Tsunami Loads and Effects

Analysis procedure using Tsunami Risk Category (TRC) Structure Classification


the Tsunami Design Zone
TRC IV
TDZ Maps
TRC II TRC III (excluding TVERS) TRC IV-(TVERS)

Energy Grade Line Required Required Required Required


Analysis (EGLA)

Permitted Permitted Required if EGLA Required


Downloaded from ascelibrary.org by La Trobe University on 09/21/20. Copyright ASCE. For personal use only; all rights reserved.

Site Specific Analysis inundation depth ≥


12 ft (3.7 m)*
TVERS: Tsunami Vertical Evacuation Refuge Shelter
* Inundation depth includes subsidence and sea level rise components
Figure G5-1. Inundation flow analysis procedure required for various tsunami risk categories.

reliability level of structural performance. If the flow depth and velocity results determined
by site-specific PTHA exceed those determined using EGLA, then the higher values must be
used in the design.
For tsunami vertical evacuation refuge structures, which are a special class of TRC IV
structure, site-specific PTHA is required regardless of the flow depth determined from EGLA
(see the exception to ASCE 7-16, Section 6.5.2).

5.3  Sea Level Change


ASCE 7-16, Section 6.5.3 requires that sea level change be considered assuming a minimum
project life cycle of 50 years. Many coastal facilities will have an anticipated life cycle
considerably longer than 50 years, so this should be taken into account when determining
the adverse effects of sea level change. At a minimum, the designer must include any sea level
rise based on the historically recorded sea level change rate for the project site, available
from the NOAA Center for Operational Oceanographic Products and Services (http://
tidesandcurrents.noaa.gov/sltrends). It might also be appropriate to consider current
research models that indicate the rate of sea level rise would be accelerating compared with
historical records (IPCC 2014). The resulting sea level rise must be added to the reference sea
level and to the tsunami runup elevation as illustrated in ASCE 7-16, Figure C6.1-2.
Any drop in sea level caused by uplift at subduction zones or other factors may not
be considered to lower the potential tsunami inundation. This is primarily because a tsu-
nami could occur at any time during the project life, and the projected uplift may not have
occurred by that time. In addition, if the tsunami is caused by a near-source subduction
earthquake, much of the historically recent uplift will be lost through co-seismic subsidence.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


6
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

Inundation Depths and Flow


Velocities Based on Runup
As described in Section 5.1, the Energy Grade Line Analysis (EGLA) provides an efficient
method for determining flow depth and velocity at any location in the tsunami design zone.
This method uses the runup elevation at the inundation limit shown in the TDZ maps to
develop a profile of the maximum flow depth and maximum flow velocity between the runup
point and the coastline. EGLA must be applied to all structures within the TDZ that require
tsunami design. For TRC IV structures, both EGLA and site-specific PTHA are required as
explained in more detail in Chapter 7.
The first step to performing an energy grade line analysis is to establish the topographic
transect profiles for the project site based on ASCE 7-16, Section 6.8.6.1, which states, “The
principal inflow direction shall be assumed to vary by +22.5 degrees from the transect per-
pendicular to the orientation of the shoreline averaged over 500 ft (152 m) to either side of
the site.”
Figure G6-1 illustrates this procedure using the ASCE Tsunami Design Geodatabase for
the Seaside, Oregon, location. The center transect is drawn perpendicular to 1,000 ft of shore-
line, centered on the project site. The geodatabase has a protractor guide that helps define the
22.5 degree angles from the center (C) transect to the clockwise (CW) and counterclockwise
(CCW) transects. One should then find the transect within this 45 degree arc that produces
the most severe flow conditions at the project site. In practice, the variation of flow conditions
is not generally that dramatic, so it is generally adequate to consider the maximum results
from the three transects shown—center, clockwise, and counterclockwise—to determine the
controlling design flow conditions. The topographic profile for each of the transects can be
obtained by drawing a transect in the ASCE Tsunami Design Geodatabase (https://asce7tsu-
nami.online/) and exporting the transect as a data file. Although Google Earth Pro will provide
an elevation profile for a section line, the elevations are not particularly accurate and should
not be relied on for EGLA. The runup elevations at which the EGLA will initiate are determined
from the runup elevation points for each transect as shown in Figure G6-1. Application of
EGLA is described in more detail in Section 6.2 of this guide.

43
@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


44 Tsunami Loads and Effects
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

Figure G6-1. Energy grade line analysis transects for Seaside, Oregon, project site.
Source: ASCE Tsunami Design Geodatabase (2016).

6.1  Maximum Inundation Depth and


Flow Velocities
Video analysis of flow velocities during past major tsunamis, most notably the 2011 Tohoku
tsunami, indicates that the potential maximum feasible flow velocities on land are on the
order of 33 ft/s (10 m/s) (Chock et al. 2013b). Because of the potential for flow amplification
due to obstructions in the built environment, as discussed in Section 6.5 of this guide, a max-
imum amplification of 1.5 is possible, leading to a limiting maximum flow velocity of 50 ft/s
(15.2 m/s). Any energy grade line analysis results that exceed this velocity can be capped at
50 ft/s (15.2 m/s). This limit can be reduced to 1.5(ghmax)1/2 if this results in a lower velocity.
This is equivalent to a Froude number of 1.5, which is well above the critical value of 1.0, which
separates subcritical from supercritical flow. Froude numbers exceeding 1.5 are very unlikely
to occur except for extremely steep terrain where the presence of structures is unlikely.
To avoid underestimation of flow velocities near the inundation limit, a minimum veloc-
ity of 10 ft/s (3.0 m/s) is prescribed, regardless of the results from EGLA. Theoretically for a
2D transect analysis such as EGLA, the flow velocity at the inundation limit is zero. However,
this ignores the 3D aspects of flow circulating parallel to the inundation limit.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Inundation Depths and Flow Velocities Based on Runup 45

During development of the EGLA methodology, a Monte Carlo simulation was performed
by applying EGLA to hundreds of randomly generated topographic transects and comparing
the results to those of a numerical inundation model (Kriebel et al. 2017). It was found that
if the runup elevation was below the maximum topographic elevation along the transect,
EGLA could be unconservative or might not converge to a solution. To retain the simplic-
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

ity of the EGLA approach while ensuring that it provided consistently conservative results,
ASCE 7-16, Section 6.6.1 requires that the runup elevation be increased to the point that will
match the highest point along the flow transect. Consider the topographic transect shown
in Figure G6-2. Assume that the runup elevation and inundation limit given in the TDZ map
were 13.9 ft and 898 ft, respectively, represented by Point A on the transect. Because there is
a location between the shoreline and this inundation limit that has a higher elevation (15.7 ft),
the runup elevation must be increased to match that higher elevation. Therefore, the runup
elevation used in EGLA will be 15.7 ft, and the analysis will start from the topographic point
matching that elevation, represented by Point B in Figure G6-2. The appropriate input values
for EGLA are therefore runup elevation of 15.7 ft and inundation limit of 1,144 ft.

Figure G6-2. Example of runup elevation modified to match highest topographic elevation between
shoreline and inundation limit (according to ASCE 7-16, Section 6.6.1).

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


46 Tsunami Loads and Effects

6.2  Energy Grade Line Analysis


EGLA provides a simplified approach to determine estimates of the maximum flow depth
and maximum flow velocity at the project site given only the information contained in the
tsunami design zone maps (Kriebel et al. 2017). The method is a stepwise procedure, starting
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

from the runup elevation at the mapped inundation limit and working shoreward to obtain
the flow parameters at the site of interest. The runup points on the TDZ maps were deter-
mined using a nonlinear inundation model, calibrated to the 2,500-year probabilistic off-
shore wave height, as described in Section 3.3 of this guide. This implies that the runup point
represents the point at which all the incoming tsunami energy has been dissipated, either
by topographic elevation change or by friction as the flow travels over the terrain. EGLA
therefore begins at the runup point with zero energy, and then reaccumulates energy due
to topographic slope and friction as one moves along the transect toward the shoreline. The
method has been extensively validated against 35,000 randomly generated transects and
multiple field survey comparisons from recent tsunamis (Kriebel et al. 2017). These com-
parisons were used to statistically bias the results of EGLA so that the resulting momentum
flux (hu2) would be approximately one standard deviation above the mean, providing some
conservatism in the resulting structural loads for achieving design reliability targets. (Refer
to Section 7.6 for an explanation of momentum flux.)

6.2.1  EGLA Theory


The theory behind EGLA is illustrated in Figure 6.6-1 in ASCE 7-16, reproduced here as Fig-
ure G6-3. The energy head, Eg,i, at any point along the transect is made up of a potential
head, hi, and a kinetic head, ui2/2g (Review G6-1). At the runup point, both flow depth and
velocity are zero, so there is no energy. As one moves back toward the shoreline, energy is
accumulated from downward slope in the topography, φiDxi, where φi is the average slope

Figure G6-3. Schematic of energy grade line analysis (ACE 7-16 Figure 6.6-1).

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Inundation Depths and Flow Velocities Based on Runup 47

Review G6-1. Energy Head.

When a fluid of depth h travels over a horizontal plane with velocity u, then the total
energy consists of both potential and kinetic energy.
2
The potential energy is given by EP = r gh
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

1
and the kinetic energy is given by EK = mu 2
2
where ρ = Fluid mass density,
g = Acceleration due to gravity,
h = Flow depth,
m = Mass of water in a column of water with unit plan area, and
u = Fluid velocity.

The mass of a water column of unit plan area and height h is m = ρh.
1
Therefore, the kinetic energy can be written as EK   hu 2 .
2
Dividing both energy terms by ρ gh gives the energy as a hydraulic head.
The potential energy head is EPh = ρgh2/ρgh = h
1
 hu 2 u 2
and the kinetic energy head is EKh 2  .
 gh 2g
Therefore, the total energy head at any point i is Eg,i = hi + ui2/2g.

of the transect between two adjacent points on the transect and Dxi is the horizontal dis-
tance between those two points. Energy is also accumulated through recovery of the friction
losses as the water flowed over the terrain, represented by siDxi, where si is the friction slope
2
given by si   ui 
 1.49  4 / 3 (in US customary units).
 hi )
2
/ (
 n 

6.2.2  EGLA Procedure


Application of EGLA will be demonstrated using the Monterey, California, site. This example
was also used to develop the example EGLA spreadsheets included in the ASCE 7-16 Com-
mentary as Figures C6.6-1 and C6.6-2. For the reader to be able to relate this example to the
values in the commentary spreadsheet, the same input parameters have been used here,
even though the final TDZ values are slightly different.
The EGLA procedure consists of the following steps as outlined in Figure C6.6-1 in the
ASCE 7-16 Commentary:

1. Obtain the runup values from the tsunami design zone map for each of the three tran-
sects as shown for the Monterey, California, site in Figure G6-4. If the inland end of a

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


48 Tsunami Loads and Effects

transect does not stop at one of the TDZ data points, the runup elevation can be interpo-
lated between the nearest two data points on the TDZ. For example, the center transect
at the Monterey building site ends approximately one-third of the distance between
two adjacent TDZ data points. Interpolation gives the runup elevation as 24.59 ft relative
to mean high water (MHW) and 29.36 ft relative to North American Vertical Datum 1988
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

(NAVD88). Note that both MHW and NAVD88 elevations are listed for the coastlines of
Washington, Oregon, and California. The difference between NAVD88 and MHW at this
location is 4.77 ft but varies along the coastline. Future editions of ASCE 7, Chapter 6 will
only include reference to the MHW elevations because they are more commonly used
for tsunami modeling. Note that the center transect end point runup differs from the
26.87 ft NAVD88 value used in the commentary spreadsheet because the final TDZ maps
were slightly different from those used to generate the spreadsheet. In this example
we will apply the spreadsheet input so that the reader can follow the steps as they are
processed in the ASCE 7-16 commentary example spreadsheet.
Using the DEM available in the ASCE Tsunami Design Geodatabase, each topo-
graphic transect is obtained as a series of x-z grid coordinates of the points defining a

Figure G6-4. Google Earth image of Monterey, California, showing the project site and the three
transects used for the energy grade line analysis. The center transect is perpendicular to the shoreline
averaged over a distance of 500 ft on either side of the center transect. This transect is then rotated
about the project location by 22.5 degrees to get the clockwise transect and –22.5 degrees to get the
counterclockwise transect.
Source: ASCE Tsunami Design Geodatabase (2016).

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Inundation Depths and Flow Velocities Based on Runup 49

series of segmented slopes, in which x is the distance inland from the shoreline to the
point and z is the ground elevation of the point.

2. The center topographic transect for the Monterey, California, example is shown in Fig-
ure G6-5. The elevations are referenced to the NAVD88 datum for the shoreline eleva-
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

tion. To ensure accuracy of EGLA, the horizontal spacing of the points on this transect
must be less than 100 ft. In Figure G6-5 the spacing is 16.4 ft, which satisfies the horizon-
tal spacing requirement. The runup elevation determined from the TDZ maps (26.87 ft)
defines the point on the transect where EGLA will begin. The corresponding inundation
distance, xR = 1,505.5 ft, is determined from the topographic transect.

3. Compute the ground slope, φi, of each segment along the transect. The ground slope
for the first segment seaward of the runup point, using data from Figure C6.6-2 in the
ASCE 7-16 Commentary, is

zi−1 − zi 26.87 − 25.73


fi = = = 0.0695
xi−1 − xi 1, 505.5 −1, 489.1

Figure G6-5. Topographic transect for energy grade line analysis for the Monterey, California, site.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


50 Tsunami Loads and Effects

4. Determine Manning’s coefficient, n, from ASCE 7-16, Table 6.6-1 for each segment based
on terrain analysis. For this example, Manning’s coefficient was taken as a constant n =
0.030 for all segments from the shoreline to the inundation limit, representing “all other
cases” in Table 6.6-1. Note that although there are housing structures along the entire
transect (Figure G6-4), this would qualify as “suburban density” and would not qualify
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

for the “urban density” listed in Table 6.6-1 with n = 0.040.

5. Compute the Froude number, Fr,i at each point on the transect using Equation (6.6-3).
For the first segment of the example seaward of the runup point, this gives

0.5
 x 
0.5
 1, 489.1 
Fr ,i  1    1.0 1    0.104
 xR   1, 505.5 

The Froude number coefficient, α, was taken as 1.0 when developing the ASCE  7-16
Commentary example spreadsheet, assuming that bore conditions will not occur at the
Monterey, California, site. ASCE 7-16, Section 6.6.4 gives the conditions for which bores
must be considered, as described in more detail in Section 6.4 of this guide. In reality,
bore conditions do exist at Monterey, so α = 1.3 should have been used in this example.

6. Start at the point of runup with a boundary condition of ER = 0.

7. Select a nominally small value of inundation depth, hR = 0.1 ft, at the runup point to avoid
a singularity at this location when applying Equation (6.6-2).

8. Calculate the hydraulic friction slope, si, using ASCE 7-16, Equation (6.6-2) for the first
segment of the example. This gives

gFr2,i 32.2  0.1042


si    0.000305
  1.49  2 1/ 3    1.49  2 1/ 3 
   hi     0.1 
 n     0.030  

9. The hydraulic energy head, Ei, can then be computed at point i using ASCE 7-16, Equa-
tion (6.6-1) as

Eg,i = Eg,i – 1 + (φi + si)Dxi = 0 + (0.0695 + 0.000305)(1,505.5 – 1,489.1) = 1.145 ft

10. The inundation depth, hi, can then be determined from the hydraulic head as

E g ,i 1.145
hi    1.14 ft
1  0.5F   2
r ,i 1  0.5  0.1042 

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Inundation Depths and Flow Velocities Based on Runup 51

11. Using the definition of the Froude number, Fr,i = u/(gh)1/2, the velocity at this location
will be
ui  Fr ,i ( gh)0.5  0.104(32.2 1.14)0.5  0.63 ft/s
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

12. Repeat this process to compute h and u at every point along the transect until reaching
the project site.

13. The resulting inundation depth, h, and flow velocity, u, at the site are then used as hmax
and umax in the tsunami loading calculations. Note that ASCE 7-16, Section 6.6.1 requires
that flow velocity not be less than 10 ft/s, but need not be taken greater than 1.5(ghmax)1/2
with an absolute maximum of 50 ft/s. In addition, if h < 3 ft (1 m), then tsunami design is
not required for TRC II and III structures.

It is evident that this stepwise calculation can be performed more easily, and more accu-
rately, using a spreadsheet as shown in ASCE 7-16 Commentary, Section C6.6.2. A software
package is also available to perform the EGLA (NumSoft Technologies, EGL version 2.0).
For the Monterey, California, example, the resulting inundation depth profile across the
entire transect is shown in Figure G6-6. To demonstrate the effect of changing Manning’s
number, EGLA was performed with all three values of n listed in ASCE  7-16, Table 6.6-1,
as well as the suggested maximum value of 0.050 referred to in ASCE  7-16 Commentary,
Section C6.6.3. The most appropriate value for the Monterey site, as indicated earlier, is n
= 0.030. Adding the flow depth to the topographic elevation produces the inundation ele-
vation profiles shown in Figure G6-7. The flow velocity profiles across the full transect are
shown in Figure G6-8. Note that the minimum flow velocity may not be less than 10 ft/s
(3.0 m/s) according to ASCE 7-16, Section 6.6.1. The maximum flow velocity need not exceed
1.5(ghmax)1/2 with an absolute maximum of 50 ft/s. For this example, the maximum flow depth
at the project site is 14.7 ft; therefore, the maximum velocity need not exceed

umax ≤ 1.5(ghmax)1/2 = 1.5(32.2 × 14.7)1/2 = 32.6 ft/s < 50 ft/s

Because umax = 18 ft/s, this upper limit does not control. Typically, the 1.5(ghmax)1/2 term
will not control for EGLA since the maximum Froude number is 1.3. However, the 50 ft/s may
control, and both limits are applicable for site-specific inundation modeling.
The hydrodynamic loads on a structure are proportional to the momentum flux (hu2) as
explained in Section 7.6.1 of this guide. However, the maximum flow velocity does not usually
occur simultaneously with the maximum flow depth. In fact, once the tsunami flow reaches
its maximum depth, the flow velocity is generally low since the flow direction is about to
change from incoming to outgoing flow. ASCE 7-16, Section 6.8.3.1 defines three load cases
that need to be considered for every structural design. Load Case 2 represents the maximum
momentum flux and corresponds to the maximum flow velocity occurring at two-thirds

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


52 Tsunami Loads and Effects

of the maximum flow depth. Therefore, the maximum momentum flux considered by the
ASCE  7-16 provisions is 2/3(hmaxumax2), which is plotted in Figure G6-9 for the Monterey,
California, example.
Because the momentum flux is proportional to the square of the flow velocity, it is rela-
tively low near the runup point but increases rapidly toward the coastline as the flow velocity
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

increases. It is obvious that structures located close to the shoreline will experience sig-
nificantly larger tsunami loads than those located further inland. It is therefore preferable
to locate critical facilities as far from the shoreline as practical. Note also that the effect
of Manning’s number is more evident in the momentum flux. Between an n value of 0.030,
which is appropriate for the suburban nature of the Monterey transect, and an n value of
0.050, the momentum flux, and therefore the hydrodynamic loads on the building, increases
by 40%. This level of conservatism is not necessary unless a higher value of Manning’s num-
ber is warranted by the terrain roughness.
It is only necessary to compute the flow parameter profiles from the runup location to
the project site. However, it may be useful to view the entire inundation depth, flow velocity,
and momentum flux profiles to understand how the tsunami flow varies with topography and
distance from the shoreline. It is also illustrative to consider variations in Manning’s n and its
effect on the predicted flow conditions at the project site, as was done for this example. Use of
a spreadsheet makes these comparisons very easy. The example shown here is identical to that
contained in the spreadsheet in ASCE 7-16 Commentary, Section C6.6.2. A software package is
also available from NumSoft Technologies to perform the energy grade line analysis.

Figure G6-6. Inundation depth (hi) profile from energy grade line analysis.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Inundation Depths and Flow Velocities Based on Runup 53
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

Figure G6-7. Inundation elevation (hi + zi) profile from energy grade line analysis.

Figure G6-8. Flow velocity (ui) profile from energy grade line analysis.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


54 Tsunami Loads and Effects

To illustrate the effect of the α-value in ASCE 7-16, Equation (6.6-3) (Step 5 of the EGLA
procedure), the Monterey, California, example was repeated with Manning’s n = 0.030 and
α = 1.0 and 1.3. Figure G6-10 shows that the inundation depth at the project site decreases
by 8%, and Figure G6-11 shows that the flow velocity increases by 24% when the α-value is
changed from 1.0 to 1.3. This is to be expected for bore conditions (α = 1.3) that have greater
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

velocity. As shown in Figure G6-12, the resulting momentum flux increases by 42%, indicat-
ing the importance of correctly selecting the α-value when using EGLA.

Figure G6-9. Maximum momentum flux corresponding to Load Case 2 (2/3hu2).

Figure G6-10. Inundation depth (hi) profile from energy grade line analysis for different α-values.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Inundation Depths and Flow Velocities Based on Runup 55
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

Figure G6-11. Flow velocity (ui) profile from energy grade line analysis for different α-values.

Figure G6-12. Maximum momentum flux corresponding to Load Case 2 (2/3hu2) for different α-values

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


56 Tsunami Loads and Effects

6.2.3  EGLA Applied to Overwashed Peninsula


When tsunami flow overtops an island or peninsula and enters a water body behind the over-
washed feature, then the energy grade line analysis approach described previously does not
apply because there are no runup points on the overwashed terrain that can be used to initi-
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

ate the EGLA procedure. For this purpose, inundation depth points were added to the ASCE
Tsunami Design Geodatabase to control the flow depth estimates generated using EGLA.
In addition, a modification to Section 6.6.1 of ASCE 7-16 was made as part of Supplement 1
issued by ASCE on December 12, 2018 (ASCE 2018). This supplement adds the following text:

Where the site lies within a completely overwashed area for which Inundation Depth
Points are provided in the ASCE Tsunami Design Geodatabase, the inundation ele-
vation profiles shall be determined using the Energy Grade Line Analysis with the
following modifications:

(a) The Energy Grade Line Analysis shall be initiated from the inland edge of the
overwashed land with an inundation elevation equal to the maximum topographic
elevation of the overwashed portion of the transect.

(b) The Froude number shall be 1 at the inland edge of the overwashed land and shall
vary linearly with distance to match the value of the Froude number determined at
the shoreline per the coefficient α.

(c) The Energy Grade Line Analysis flow elevation profile shall be uniformly adjusted
with a vertical offset such that the computed inundation depth at the Inundation
Depth Point is at least the depth specified by the ASCE Tsunami Design Geodatabase,
but the flow elevation profile shall not be adjusted lower than the topographic eleva-
tions of the overwashed land transect.

An example of this procedure is demonstrated using a site in Oysterville on the over-


washed Long Beach Peninsula, Washington State (Figure G6-13). A transect is cut through
the project site and perpendicular to the ocean shoreline. Two additional transects at
±22.5 degrees would also need to be cut but will not be shown here for simplicity. The tran-
sect crosses the inundation depth point line at a location with ground elevation of 5.69 ft and
inundation depth of 37.81 ft (Figure G6-13). The transect obtained from the ASCE Tsunami
Design Geodatabase is shown as the black solid line in Figure G6-14, along with the inunda-
tion depth point location (at 5.69 ft elevation) and the project site (at 11 ft elevation).
The energy grade line analysis is initiated from the inland side of the peninsula with an
initial flow depth equal to the highest topographic point along the transect. This is 50.2 ft
for this particular transect. The Froude number is assumed to be 1.0 at this location. Bores

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Inundation Depths and Flow Velocities Based on Runup 57
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

Figure G6-13. ASCE Tsunami Design Geodatabase image showing Oysterville site on overwashed Long
Beach Peninsula, Washington.
Source: ASCE Tsunami Design Geodatabase (2016).

Figure G6-14. Energy grade line analysis applied to overwashed peninsula.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


58 Tsunami Loads and Effects

must be considered at the shoreline because of the low slope bathymetry offshore from the
project site (ASCE 7-16, Section 6.6.4). Therefore, the Froude number at the shoreline is 1.3
and varies linearly to 1.0 across the overwashed peninsula.
The results of this initial EGLA are shown as the green dashed line in Figure G6-14. The
flow depth at the inundation control point must be compared with the flow depth of 37.81 ft
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

provided by the ASCE Tsunami Design Geodatabase. As seen in Figure G6-14, the initial EGLA
gives a flow depth of 60.67 ft at this location. Therefore, the entire flow depth profile can
be lowered by 60.67–37.81 = 22.86 ft to get the final EGLA flow elevation shown by the blue
solid line in Figure G6-14. If any point along this flow elevation line drops below the ground
elevation, then the flow line cannot be lowered any further, regardless of the inundation
control point value. This does not occur in this example so the value for hmax at the project
site is 27.7 ft as shown in Figure G6-14. The flow velocities determined using the initial EGLA
are shown in Figure G6-15. They remain unchanged when the flow depth profile is raised or
lowered to match the inundation control point. The maximum flow velocity, umax, at this site
would therefore be 46.7 ft/s.
For this example, the initial flow depth exceeded the value at the inundation control
point, so the flow depth line could be lowered. However, if the initial flow depth line fell
below the inundation control point, then the line would have to be increased until the depth
at the control point matches the value given in the ASCE Tsunami Design Geodatabase.
This modification of the EGLA method for overwashed peninsulas has been incorpo-
rated into Version 2.0 of the EGL software (https://www.numsofttechnologies.com/) pro-
duced by NumSoft Technologies.

6.3  Terrain Roughness


The TDZ maps introduced in ASCE 7-16, Section 6.1 are based on inundation modeling using
bare-earth assumptions. In other words, the presence of vegetation, buildings, and other
structures is ignored, and the ground roughness is modeled with a Manning’s coefficient, n =
0.03, for the nearshore bathymetry and the on land topography (Review G6-2). This is a con-
servative assumption since the presence of obstructions would generally reduce the inunda-
tion limit and runup elevation. When applying the energy grade line analysis, it is no longer
conservative to use a low Manning’s roughness, as this will result in less energy required to
cause the mapped inundation. The Manning’s roughness should be selected from ASCE 7-16,
Table 6.6-1 based on the on-land conditions in the TDZ. When in doubt, a larger Manning’s
roughness will generally result in more conservative flow depths and velocities based on the
energy grade line analysis. The Manning’s roughness can vary for each segment of the EGLA
transect to represent the appropriate roughness for the terrain in that segment. It might be
appropriate to select Manning’s roughness values that reflect potential future development.
Further discussion of the Manning’s roughness is provided in ASCE 7-16 Commentary, Sec-
tion C6.6.3.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Inundation Depths and Flow Velocities Based on Runup 59
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

Figure G6-15. Flow velocity profile for overwashed peninsula example.

Figure G6-16. Determination of prevailing nearshore bathymetric slope for Monterey, California,
project location.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


60 Tsunami Loads and Effects

Review G6-2. Manning’s Roughness.

Manning’s equation is an empirical formula used to estimate average velocity for gravi-
ty flow, such as open channel flow. Because the velocity depends in part on the rough-
ness of the channel, an empirically determined roughness coefficient, n, is included in
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

the formula, given by

 A
2/3

1.49   s1/ 2
u P in US customary units
n

 A
2/3

  s1/ 2
P
u  in SI units
n
where
u = Average cross-sectional velocity,
A = Cross-sectional area of flow,
P = Wetted perimeter, and
s = Slope of the hydraulic grade line, which equals the channel bed slope when
the water depth is constant.

Values of n used in open channel flow vary from 0.0125 for smooth concrete to
0.050 for rough stone lining.
For an infinitely wide channel, representing tsunami inundation on a coastal plain,
A/P = h, where h is the depth of flow. Substituting into the preceding equation gives

1.49h 2 / 3 s1/ 2
u= in US in-lb units
n
2
(u )
Solving for s gives s= in US in-lb units
 1.49 2 4 /3
 
 n  h

which is ASCE 7-16, Equation (6.6-2).


Typical Manning’s roughness coefficients for tsunami inundation are given in
ASCE 7-16, Table 6.6-1. ASCE 7-16 Commentary, Section C6.6.3 also provides an expres-
sion to determine n for built environments with predominantly concrete, masonry,
and structural steel buildings. However, when EGLA is used with values of n greater
than 0.050, the results are significantly more conservative than results predicted by
more sophisticated numerical models.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Inundation Depths and Flow Velocities Based on Runup 61

Review G6-3. Bore Formation.

A tsunami bore can be viewed as a traveling hydraulic jump. A conventional stand-


ing hydraulic jump forms when fast-moving or supercritical water enters a volume of
slower-moving water, such as the supercritical flow down a dam spillway entering a
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

stilling basin. A stationary hydraulic jump will form as the supercritical flow transforms
to subcritical flow. The Froude number, Fr, is a dimensionless value that describes
whether flow is subcritical or supercritical as follows:

u
Fr =
gh
where
u = Water velocity,
g = Acceleration due to gravity, and
h = Flow depth.

If Fr > 1, then the flow is supercritical, and if Fr < 1, then the flow is subcritical.
If the incoming tsunami flow is supercritical and it passes over the top of stand-
ing water in a coastal lagoon or on land from a preceding wave, then a bore is likely to
form. When the Froude number of the tsunami flow drops below 1.0, then the bore will
dissipate, and the tsunami will continue as a surge.

6.4  Tsunami Bores


Tsunami bores often form as a tsunami wave approaches land (Review G6-3). A bore would
initiate when the wave height reaches about 80% of the water depth, leading to wave break-
ing and bore formation. This condition occurs most commonly when the nearshore bathy-
metric slope is milder than 1/100, or when there is a shallow fringing reef or similar step
discontinuity in the nearshore bathymetry (Madsen et al. 2008, Murata et al. 2010). Bore con-
ditions must also be considered if historical records, recognized literature, or site-specific
modeling indicates that bores may form at the site. Example G6-1 shows the evaluation of
bore conditions at the four prototypical building sites introduced in Chapter 2 of this guide.
Because of the high velocity and steep leading edge of a turbulent bore compared with
a tsunami surge, larger forces can develop on certain structural elements. If tsunami bores
need to be considered at a particular site, then the provisions of ASCE 7-16, Section 6.10.2.3,
“Tsunami Loads on Vertical Structural Components,” particularly walls, and ASCE 7-16, Sec-
tion  6.10.3.3, “Tsunami Bore Flow Entrapped in Structural Wall-Slab Recesses,” must be
applied. In addition, as described in ASCE 7-16, Section 6.2, α = 1.3 is used in EGLA for bore
conditions, leading to higher flow velocities and momentum flux.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


62 Tsunami Loads and Effects

Example G6-1. Tsunami Bore Occurrence Example.

ASCE 7, Section 6.6.4 requires that “tsunami bores shall be considered where any of the
following conditions exist:
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

1. The prevailing nearshore bathymetric slope is 1/100 or milder,


2. Shallow fringing reefs or other similar step discontinuities in nearshore
bathymetric slope occur,
3. Where historically documented,
4. As described in the recognized literature, or
5. As determined by a site-specific inundation analysis.”

The “prevailing nearshore bathymetric slope” can be determined from the 2,500-
year offshore tsunami wave conditions since they are all documented at 100 m (328 ft)
bathymetric depth. Using the measure tool in the ASCE 7 Tsunami Geodatabase to
determine the distance, x100 from the 100 m bathymetric line to the shoreline gives the
nearshore slope as φ = 100/x100 in SI units or φ = 328/x100 in US customary units.
The following table shows that for the four sites selected for the prototypical
building studies (Table G2-1), tsunami bores must be considered at each location. Fig-
ure G6-16 shows how the nearshore slope value was obtained for Monterey, California.

Nearshore Fringing Historical Recognized Inundation Consider


Location Slope Reef Bores Literature Analysis Bores
Hilo, HI 1/84 No Yes Yes N/A Yes
Seaside, OR 1/250 No No N/A N/A Yes
Monterey, CA 1/123 No No N/A N/A Yes
Waikiki, HI 1/26 Yes No N/A N/A Yes
Notes: N/A = not available. Red numbers = factor that requires consideration of bores.

6.5  Amplified Flow Velocities


Flow velocities determined using the energy grade line analysis are based on a 2D transect
cut perpendicular to the shoreline. They therefore ignore any 3D effects of nonuniformity of
the transect as one moves along the shoreline. There is also no consideration for focusing of
flow around and between sturdy structures that do not collapse. For this reason, it is neces-
sary that the flow velocities determined using EGLA be amplified by one of the approaches
provided in ASCE 7-16, Section 6.8.5. The amplification approaches are addressed in greater
detail in Section 8.5 of this guide. The amplified flow velocity need not exceed the maximum
limit of 50 ft/s (15.2 m/s) specified in ASCE 7-16, Section 6.6.1.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


7
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

Site-Specific Tsunami Inundation

ASCE 7-16, Section 6.5 requires that site-specific tsunami inundation modeling be performed
for all TRC IV buildings where EGLA determines that the inundation depth at any point within
the footprint of the structure exceeds 12 ft (3.66 m). In addition, site-specific inundation
modeling is required for all tsunami vertical evacuation refuge structures regardless of the
inundation depth at the site. Site-specific tsunami inundation modeling may also be used
for TRC III and II buildings but is not required. A detailed knowledge of tsunami modeling,
propagation, and inundation is required to perform a reliable site-specific tsunami inunda-
tion analysis. Models used for this purpose must comply with the benchmarking verification
standards listed in ASCE 7-16, Section 6.7.6.7.
Whenever a site-specific tsunami inundation analysis is performed, the resulting max-
imum flow velocity at the project site must be compared with the maximum velocity deter-
mined using the energy grade line analysis for the same site. In urban environments, where
Manning’s coefficient of n = 0.040 is used in the EGLA, the flow velocities at the project site
may not be less than 90% of those determined from EGLA before any velocity adjustments
caused by flow amplification. For other terrain roughness conditions, the limit is reduced to
75% (see ASCE 7-16, Section 6.7.6.8). This check provides some protection against potential
model inaccuracies and errors when performing the site-specific inundation modeling.
ASCE  7-16 Commentary, Section C6.7 provides the methodology for performing a site-
specific PTHA that is generally consistent with probabilistic seismic hazard analysis (PSHA)
used to develop seismic design provisions in ASCE 7-16. The intent of the PTHA is to determine
the time history of flow parameters (depth, velocity, and momentum flux) at the project site for
the specified 2,500-year return period, incorporating allowances for both aleatory and epis-
temic uncertainties. Briefly, aleatory uncertainty can be thought of as the random variability in
results when the same experiment is performed multiple times, whereas epistemic uncertainty
is the result of effects that are imperfectly characterized or ignored in the modeling analysis.
In the ASCE 7-16 PTHA, a statistically weighted logic tree approach was used to account for the
epistemic uncertainties in the tectonic parameters, and a probability distribution was used to
account for the aleatory uncertainties evidenced in the inundation results compared to obser-
vations when real tsunamis are analyzed with the model.

63
@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


64 Tsunami Loads and Effects

Considering all possible seismic sources, tsunami waveforms are generated and prop-
agated in deep water using linear long-wave equations, combined in a linear fashion at the
328 ft (100 m) depth contour, to determine the highest offshore wave heights and to develop
period and waveform parameters that represent the 2,500-year tsunami event. This pro-
cess was used to generate the offshore tsunami amplitude, HT, and predominant period,
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

TTSU, that constitute the ASCE database of geocoded reference points available at https://
asce7tsunami.online/ (Thio et al. 2017). ASCE 7-16, Figure C6.7-1 shows a sample of the off-
shore tsunami amplitude database contained in the ASCE Tsunami Design Geodatabase.
Clicking on any of the blue squares at the 328 ft (100 m) offshore bathymetric line will open
a pop-up window with tsunami amplitude, predominant period, and the longitude and lati-
tude of the point. Also provided is an image of the disaggregated sources contributing to the
2,500-year tsunami amplitude at that location. These data are also available as a data file for
use in other mapping software such as ArcGIS and MATLAB, among others. These data cover
the entire coastlines of Washington, Oregon, California, and Hawai‘i, and the majority of the
tsunami-threatened southern coastline of Alaska, as shown in ASCE 7-16, Figure 6.7-1.
The 328 ft (100 m) offshore bathymetric contour was selected as the reference offshore
wave location because tsunami waves still behave linearly at this depth. This greatly sim-
plifies the probabilistic combination of different tsunamigenic sources to develop the tsu-
nami waveform. It also removes the complexity of nearshore bathymetric effects of shoaling,
focusing, resonance, dispersion, and bore formation from the required comparison between
a hazard-consistent tsunami-generated waveform and the ASCE 7-16 specified waveform.
To perform a site-specific tsunami inundation analysis, it is not necessary to repeat the
PTHA. One can apply either of two approaches to propagate the tsunami on land to deter-
mine the inundation zone and runup limits, along with time histories of flow depth, velocity,
and momentum flux at the location of interest. The first approach is to start with the tsunami
waveform specified in ASCE  7-16, Section 6.7.2 at the 328 ft (100 m) offshore bathymetric
line and propagate that waveform on land. The second approach is to use the disaggre-
gated sources generated by the ASCE PTHA (also available at https://asce7tsunami.online/)
to regenerate tsunami waves at the various sources, propagate them across open ocean, and
model the inundation as a continuous process. This procedure is usually performed using a
series of nested grids with increasingly fine resolution as one nears the coastline of interest.
It is required that the offshore wave heights generated by this second approach at the 328 ft
(100 m) offshore water depth be compared with those in the ASCE Tsunami Design Geoda-
tabase, with no single point falling below 80% of the corresponding mapped value. Although
not specifically required in ASCE  7-16, it is also the intent that the offshore wave heights
match on average. If the generated wave heights are less than 80% of the values given in the
ASCE Tsunami Design Geodatabase, then the magnitude of the source mechanism must be
increased until this minimum threshold is reached along the shoreline of interest. Because
the offshore wave heights include probabilistic estimates of the aleatory and epistemic
uncertainties, the seismic magnitude at a particular source required to match these wave
heights may be larger than historical earthquakes at that source. This second approach was

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Site-Specific Tsunami Inundation 65

used to generate the tsunami design zone maps and runup elevations available at https://
asce7tsunami.online/.

7.1  Tsunami Waveform


Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

If one elects to generate the site-specific on land inundation by starting with the 328 ft
(100 m) offshore wave condition, then ASCE 7-16, Equation (6.7.1-1) provides the waveform
that is to be used. This waveform consists of a crest with amplitude equal to the tsunami
amplitude and a trough with amplitude between 43% and 100% of the tsunami ampli-
tude, depending on the region, as specified in Table 6.7-1 and illustrated in Figure 6.7-2 of
ASCE 7-16. The wave period, TTSU, is obtained from the ASCE Tsunami Design Geodatabase.
Only in regions not covered by the ASCE Tsunami Design Geodatabase may the TTSU values
provided in Table 6.7-1 be used.
Two possible waveforms must be considered based on the maximum and minimum val-
ues listed for a2 in ASCE 7-16, Table 6.7-1. These specified waveforms must each be applied
twice, once with a leading crest and once with a leading trough, to simulate both possible
tsunami wave conditions. Example G7-1 shows development of the tsunami waveforms for
the prototypical building site in Monterey, California. Note that ASCE 7-16, Table 6.7-1 refers
to Northern California and Southern California. For tsunami modeling purposes the dividing
line between these two is Cape Mendocino.
If one elects to use an integrated generation, propagation, and inundation model from
the disaggregated tsunami sources that matches the offshore wave amplitudes as required
by ASCE 7-16, Section 6.7.5.1.4 (i.e., not less than 80% of the corresponding ASCE Tsunami
Design Geodatabase amplitudes), then the values given in Table 6.7-1 need not be used.

Example G7-1. Example Waveform Development for Monterey, California.

Figure G7-1 shows an image of Monterey, California, with the ASCE 7-16 offshore wave
parameters from the ASCE Tsunami Design Geodatabase. The parameters required
to develop the offshore waveform are the amplitude, HT = 11 ft, and the period, TTSU =
36 min = 2,160 s. Because Monterey is south of Cape Mendocino, the range of negative
wave amplitudes from ASCE 7-16, Table 6.7-1 are −0.43 to −0.67HT. Plotting ASCE 7-16,
Equation (6.7-1) with both of these values gives the normalized plots shown in Fig-
ure G7-2. Applying the absolute values of HT and TTSU produces the actual value plots
shown in Figure G7-3. These waveforms must be propagated toward the shoreline
with both the positive and negative pulses leading. This therefore requires four sep-
arate inundation studies to determine the maximum flow depth, flow velocity, and
momentum flux at the project site.
Recent studies have shown that it may be necessary to use a tsunami sequence
with multiple waveforms to accurately reproduce the 2,500 year tsunami inundation.
Modifications to this approach are proposed for ASCE 7-22.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


66 Tsunami Loads and Effects
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

Figure G7-1. Offshore wave parameters for Monterey, California.


Source: ASCE Tsunami Design Geodatabase (2016).

Figure G7-2. Offshore waveform at 328 ft (100 m) bathymetric depth: normalized values.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Site-Specific Tsunami Inundation 67
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

Figure G7-3. Offshore waveform at 328 ft (100 m) bathymetric depth: actual values.

7.2  Tsunamigenic Sources


If an engineer elects not to utilize the offshore wave parameters or the disaggregated sources
specified in the ASCE Tsunami Geodatabase (https://asce7tsunami.online/), then he or she
will need to perform a full PTHA based on the methodology prescribed in this section. All
possible tsunamigenic sources, including local and distant subduction zones, local fault
sources, and landslide sources, must be considered. It is not expected that this approach will
be used for regions where the ASCE offshore wave data are available, but this approach will
be useful for communities that have not been mapped or for US territories and other coun-
tries wishing to use the ASCE 7-16 tsunami design provisions.

7.3  Earthquake Rupture Unit Source


Tsunami Functions
This section describes the approach that must be used if one represents the earthquake rup-
ture slip by a linear combination of unit source functions. This involves generating Green’s
functions for the offshore location of interest for each subfault in the subduction zone being
considered, and then using a reverse analysis to estimate the required slip on each subfault
to produce the offshore wave amplitudes given by the ASCE Tsunami Design Geodatabase.
This provides a first trial seismic source that is then used to generate a tsunami and test
whether it complies with the required match at the offshore wave amplitude points. Some
adjustment of the source subfault slips may be required to produce an acceptable offshore

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


68 Tsunami Loads and Effects

wave amplitude match. The final source scenario is then considered hazard consistent with
the ASCE 7-16 2,500-year event and can be used to perform the site-specific tsunami inun-
dation modeling. This is the same approach that was used to develop the offshore wave
parameters and tsunami design zone provided in the ASCE Tsunami Design Geodatabase.
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

7.4  Treatment of Modeling and


Natural Uncertainties
This section outlines the required consideration of epistemic and aleatory uncertainties in
the source modeling, tidal elevation, and tsunami modeling used to perform the PTHA. These
uncertainties have already been included in the offshore amplitudes provided in the ASCE
Tsunami Design Geodatabase. Therefore, if one is performing a site-specific inundation
analysis following the approach in Section 7.3, then these uncertainties are automatically
included. However, if one is performing a PTHA for a location not covered by the ASCE Tsu-
nami Design Geodatabase, or for a different tsunami return period, then these uncertainties
must be included in the PTHA.

7.5  Offshore Tsunami Amplitude


This section provides alternate approaches to determining the probabilistic offshore wave
heights representing the design exceedance rate of the 2,500-year maximum considered
tsunami. Regardless of the approach used, if the location being studied is included in the
ASCE Tsunami Design Geodatabase, then the resulting offshore wave amplitudes at the
328 ft (100 m) bathymetric depth contour should, on average, match those provided in the
geodatabase, with no point lower than 80% of the wave amplitudes given in the database.

7.6  Procedures for Determining Tsunami


Inundation and Runup
This section provides the requirements for modeling tsunami inundation. Modeling of tsu-
nami wave inundation on land is significantly more complex than modeling wave propaga-
tion in the open ocean. The interaction of the tsunami wave with the nearshore bathymetry
and on-land topography, including the built environment, results in nonlinear wave dynam-
ics such as soliton fission, bore formation, flow focusing, significant cross-shore variation,
and even eddy formation. Soliton fission is caused by amplitude dispersion that results as a
tsunami wave crosses shallow bathymetry. The result is a series of waves approaching the
shoreline as observed during the Tohoku tsunami inundation of the Sendai, Japan, coastline.
Tsunami bores also form because of shallow bathymetry such as fringing reefs, or because of
gradually sloping nearshore bathymetry that forces nonlinear “breaking” of the leading edge
of the tsunami flow. The dramatic change in flow depth over a short distance is evidenced by

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Site-Specific Tsunami Inundation 69

a turbulent leading edge, which is typical of bore formation. This turbulence acts as an energy
dissipation mechanism, while the dramatic change in flow depth can lead to increased struc-
tural loads as discussed in Chapter 10. Flow focusing occurs as a tsunami wave enters a bay
or estuary that reduces in width closer to shore. Wave amplitudes and velocities increase to
accommodate passage through the narrowing bathymetry. Cross-shore variation in tsunami
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

flow is the result of the complex three-dimensional geometry of the coastline. These cross-
shore effects must be considered in site-specific inundation modeling, requiring that the
model use a 2D DEM. One-dimensional (single-transect) tsunami modeling is not adequate
to capture these effects. Eddy formation is common when tsunami waves pass obstructions
such as the ends of breakwaters or piers. They are more commonly observed during draw-
down and can have an effect on coastal structures and vessels.
Because of the complexity of nonlinear inundation modeling, ASCE  7-16 only permits
the use of tsunami inundation models that have been validated following the procedures in
ASCE 7-16, Section 6.7.6.7. A number of such models have been developed by various research-
ers but should only be used by those with a thorough understanding of fluid mechanics and
the effects of local bathymetric and topographical features on the tsunami flow.

7.6.1  Representative Design


Inundation Parameters
The site-specific probabilistic tsunami modeling will result in complete time histories of
flow depth, flow velocity, and momentum flux for the duration of the tsunami event. This will
typically include more than one tsunami wave. For structural design purposes, the intent of
the tsunami inundation modeling is to determine the flow parameters at the project site that
govern the fluid loading on the structure and the associated scour. These include the max-
imum inundation depth, maximum flow velocity, and maximum momentum flux at the site.
Note that the maximum flow velocity must be compared with that obtained using the energy
grade line analysis as described in ASCE  7-16, Section 6.7.6.8. It is also necessary for the
designer to consider all three load cases defined in ASCE 7-16, Section 6.8.3.1. In addition,
the flow conditions corresponding to the maximum momentum flux determined from the
site-specific modeling must also be considered (Review G7-1).

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


70 Tsunami Loads and Effects

Review G7-1. Momentum Flux.

When fluid flow encounters a rigid object, the stagnation pressure that develops at the
leading edge of the object is given by:
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

1 2
Pstagnation = ru
2

where ρ is the fluid mass density, and u is the fluid velocity.


The hydrodynamic force applied to the object is therefore related to this pressure
applied to the projected area of the object, Aproj = B × h, where B is the object width and
h is the flow depth or the height of the object, whichever is smaller. For example, the
hydrodynamic drag on a vertical column would be given by

1 2
Fd = r u Cd Bh
2

where Cd accounts for the object shape and is referred to as the drag coefficient.
The hydrodynamic drag can be rewritten as

1
Fd = rC B(hu 2 )
2 d

where the term hu2 is often referred to as the “specific momentum flux,” or simply as
momentum flux.
Note that the maximum momentum flux during tsunami inundation is not the
product of maximum inundation depth and maximum velocity squared, because these
two maxima do not occur at the same time. The maximum inundation depth at a par-
ticular location generally occurs right before the tsunami flow is about to recede. The
flow velocity associated with maximum inundation depth is therefore generally quite
low. ASCE 7-16, Load Case 2 represents the maximum momentum flux based on the
maximum flow velocity and two-thirds of the maximum flow depth.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


8
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

Structural Design Procedures


for Tsunami Effects
Tsunami loads consist of sustained hydrostatic and hydrodynamic loads applied to the over-
all lateral force–resisting system and to individual submerged structural elements, as well as
short-duration impact loads from floating or tumbling debris driven by both the incoming
and outgoing tsunami flow. These loads need to be resisted by the structural elements and
foundation system, which may have been compromised by scour resulting from prior tsu-
nami waves. Although the dynamic effects of seismic loading are markedly different from
tsunami loads, as explained in ASCE 7-16 Commentary, Section C6.8, the strength and ductil-
ity required for seismic design significantly improves the building capacity to resist tsunami
loads (Chock et al. 2013b).
This chapter introduces the tsunami structural performance objectives for each cate-
gory of building and other structure, as shown in ASCE 7-16 Commentary, Table C6.8-1 and
repeated here as Table G8-1. It also describes the development of design time histories of
inundation depth and flow velocity that are used to determine tsunami flow conditions at the
three load cases required for all designs.
Load combinations, importance factors, and acceptance criteria for system and compo-
nent design are described. The effects of flow velocity amplification and flow directionality

Table G8-1. Tsunami Performance Levels per Tsunami Risk Category.

Tsunami performance level objective

Hazard level and tsunami Collapse


frequency Immediate occupancy Damage control prevention
Maximum considered Tsunami vertical TRC IV and TRC III TRC III and TRC II
tsunami evacuation refuge critical facilities
(2,500-year mean structures
recurrence interval)
Source: ASCE 7-16, Table C6.8-1.

71
@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


72 Tsunami Loads and Effects

are described and demonstrated by examples. Provisions are also provided for physical mod-
eling of tsunami flow, loads, and effects for those interested in performing laboratory inves-
tigations to augment the code provisions.

8.1  Performance of Tsunami Risk Category II and


Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

III Buildings and Other Structures


TRC II buildings of substantial height and TRC III buildings could have occupants at the time
a tsunami occurs, making it important from a life safety perspective that these buildings do
not collapse during the tsunami. They are therefore required to be designed for a collapse
prevention performance level. TRC III buildings such as schools with mass public assembly
occupancies and larger health care facilities with 50 or more resident patients may not be
able to evacuate in time, even though evacuation planning should be in place as the preferred
response to a tsunami warning. TRC II buildings of substantial height may also be used as
a “refuge of last resort” if people find that they have insufficient time to follow established
evacuation plans to high ground or designated vertical evacuation refuges.
During the Tohoku tsunami, tens of thousands of refugees survived by evacuating verti-
cally into buildings of adequate height and strength, even though they were not designated
as vertical evacuation buildings (Fraser et al. 2012). Similar behavior was observed during
the 2004 Indian Ocean tsunami. These “refuge of last resort” buildings have saved an enor-
mous number of lives, and there is every reason to believe that similar behavior will occur
during future tsunamis in Japan and elsewhere. At the same time, many people caught in
the inundation zone sought refuge in buildings that were either too low and were therefore
overtopped, or did not have enough structural integrity to resist the tsunami loads. The
number of people who died in this way is impossible to assess, but this loss of life can be
reduced through improved public awareness and education and by designing all buildings of
adequate height to survive the tsunami without collapse.
ASCE 7-16 Commentary, Section C6.8.1 concludes that “Based on analysis of prototypical
buildings, a threshold of 65 ft (19.8 m) above grade plane is deemed the height appropriate
for both reliable life safety and reasonable economic cost for TRC II buildings to resist tsu-
nami loads. Importantly, when evacuation from the inundation zone is not possible prior to
tsunami arrival due to long evacuation distances, road congestion, or damaged infrastruc-
ture, the public will attempt to use taller buildings to escape the tsunami inundation and
will inherently expect that such taller structures will not collapse during the tsunami.” It is
critical that emergency planners and local jurisdictions have enough buildings of adequate
height that are designed for tsunami loads, including those in TRC II, and that the public is
aware of the height required before a building can be used as a refuge of last resort. Local
jurisdictions are encouraged to require tsunami design for TRC II buildings and to select a
suitable minimum height for these buildings based on the anticipated inundation depths in
their area of influence (Carden et al. 2016, Chock et al. 2018).

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Structural Design Procedures for Tsunami Effects 73

8.2  Performance of Tsunami Risk Category III


Critical Buildings and TRC IV Buildings and
Other Structures
The local jurisdiction may elect to designate certain Tsunami Risk Category III buildings as
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

critical facilities that perform a post-tsunami function that is critical to response and/or
recovery. These TRC III critical facilities and all TRC IV buildings and other structures must
be designed for a higher performance level than required for TRC II and other TRC III build-
ings. This performance level is referred to as “Damage Control,” and lies between “Collapse
Prevention” and “Immediate Occupancy.”
The most important TRC IV buildings are tsunami vertical evacuation refuge structures
because people will be instructed to seek refuge in these structures and should therefore
have an increased assurance that their lives will not be in jeopardy. The building structure
must meet immediate occupancy performance, while nonstructural components at the ref-
uge levels must also satisfy immediate occupancy performance following the provisions in
Section 6.14 of ASCE 7-16.
For other TRC IV and Critical TRC III buildings, an intermediate performance level of
damage control, which falls between immediate occupancy and collapse prevention, is tar-
geted by the provisions in Section 6.8.2 of ASCE  7-16. These provisions include elevation
of operational nonstructural components and equipment necessary for essential functions
above the inundation elevation for the MCT. The structural components and connections,
including foundations, must be designed for damage control. Although the overall facility
may not meet the requirements of immediate occupancy, it will experience limited structural
damage and will facilitate rapid reinstatement of full operational function after the tsunami.

8.3  Structural Performance Evaluation


This section defines the structural performance evaluation criteria for application in the
design of buildings and other structures for tsunami loads. After defining load cases, impor-
tance factors, and load combinations, this section provides two acceptance criteria to eval-
uate the anticipated structural performance.

8.3.1  Load Cases


Three specific load cases are defined in ASCE 7-16, Section 6.8.3.1 for design of buildings and
other structures in the TDZ. Other load cases may also have to be considered depending on
the flow depths appropriate for particular loading evaluations, but the three load cases spec-
ified in this section must be considered as a minimum.
Load Case 1 (LC1) is intended to prevent uplift or overturning of the structure caused
by the effects of buoyancy. This load case need not be applied to open structures as defined

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


74 Tsunami Loads and Effects

in ASCE 7-16, Section 6.2 because the open nature of these buildings, or the use of tsunami
breakaway walls, allows water to flow freely through the structure. This load case will also
not apply if hydrostatic pressure cannot develop below the ground-floor slab because of
the impervious nature of the subgrade. However, typical grade-level construction calls for
a layer of gravel or sand directly below the slab, which could allow hydrostatic pressure to
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

rise with the inundation level, regardless of the permeability of the foundation soil below the
sand layer. However, this uplift pressure on the soffit of the grade-level slab will only lead to
uplift of the overall structure if the slab-on-grade is structurally connected to grade beams
and/or the foundations. If there are no grade beams, or if the slab-on-grade is isolated from
the grade beams, columns, and foundations, then the uplift pressure below the slab will only
lead to slab uplift and failure, allowing water into the interior of the building. Overall uplift of
the structure will no longer be a concern. Figure G8-1 shows two cases of structural failure
attributed in part to buoyancy where a layer of gravel or stones was present below an inte-
grated structural slab-beam system at the grade level.
Based on past tsunami damage observations, it is evident that regular window glazing
will break by the time the water level reaches the top of the window (see also Example G9-1).
Therefore, the top of the first-story windows is set as the minimum inundation depth to be
considered for buoyancy calculations. If there are no windows in the first story, then the
inundation depth must be assumed to be the full height of the first floor. Door openings
are not considered to provide ingress for water because they are likely to be closed and
have significantly greater strength than window glazing. In either case, the water depth con-
sidered for buoyancy need not exceed the maximum inundation depth determined for the
building location. Having established the inundation depth for the buoyancy calculation, the

Figure G8-1. Overturned reinforced concrete buildings in Onagawa, Japan, where failure is attributed
in part to buoyancy. Note the relatively small door or window openings at the ground-floor level
(except for an opening in the wall created by rescue crews), and structural beam-slab grade level floor
system over gravel material: (left) N 38.44072; E 141.44648; (right) N 38.44512; E 141.44683.
Source: Chock et al. (2013b).

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Structural Design Procedures for Tsunami Effects 75

corresponding flow velocity is determined using ASCE 7-16, Figure 6.8-1 so as to determine


the lateral hydrodynamic load that must be applied in conjunction with the buoyancy.
Once the governing water depth has been determined for each load condition, the total
uplift force applied to the building, Fv, owing to buoyancy is given by

Fv = gsVw (6.9-1)
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

where gs = ksgsw is the weight density of seawater including sediment effect, and Vw is the total
volume of water displaced by the building.
Example G9-3 demonstrates application of Load Case 1 to a reinforced concrete frame
building with a 1-story basement.
Load Cases 2 and 3 are best visualized using the normalized inundation depth and flow
velocity time histories shown in ASCE 7-16, Figure 6.8-1 and repeated here as Figure G8-2.
These idealized time histories were based on analysis of survivor videos taken during the
Tohoku tsunami. Apart from some high flow velocities associated with the leading edge of
the tsunami, when the flow depth is still small, the most critical time for maximum lateral
loading on the overall building was typically at a flow depth of approximately two-thirds of
the maximum, when the velocity peaks. This is specified by ASCE 7-16 as Load Case 2 (LC2)
and is assumed to represent the maximum momentum flux (hu2) and, therefore, the maxi-
mum drag on the overall structure and individual structural members.
As the flow depth reaches the maximum inundation depth, the flow velocity decreases
before reversing and starting to recede. ASCE 7-16, Load Case 3 (LC3) requires that a flow
velocity of one-third the maximum be considered to occur when the flow depth is at its max-
imum level. Although Load Case 3 is unlikely to represent the maximum lateral tsunami load
on the overall structure, it is required to ensure that structural elements within the upper
third of the maximum inundation depth are designed for appropriate hydrodynamic loads.
Figure G8-2 can also be used to determine the appropriate flow velocity at other flow depths
of interest.
If a site-specific tsunami analysis is performed in accordance with ASCE  7-16, Sec-
tion 6.7, then the resulting time histories of flow depth and velocity can be used to evaluate
the maximum hydrodynamic loads on individual members and on the overall structural lat-
eral force–resisting system. This may require the use of more than just the three load cases
listed previously. It is then not necessary to use Figure G8-2.

8.3.2  Importance Factors


Tsunami importance factors are specified in ASCE 7-16, Table 6.8-1 as 1.0 for TRC II structures
and 1.25 for TRC III and IV structures including vertical evacuation refuge structures. These
importance factors were determined by means of a comprehensive reliability analysis of the
anticipated performance of typical reinforced concrete and structural steel load-bearing
columns in prototypical buildings (Chock et al. 2016). The results of this reliability analysis

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


76 Tsunami Loads and Effects
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

Figure G8-2. Normalized tsunami inundation depth and flow velocity time histories.
Source: ASCE 7-16, Figure 6.8-1.

are compared with the anticipated reliabilities for structures designed to ASCE 7-16 seismic
provisions in Table G8-2.

For TRC II structures, this reliability analysis indicated that an importance factor of 1.0
would result in a 7.5% probability of the tsunami hydrodynamic loads exceeding the column
capacity during the maximum considered tsunami. This can be compared with an antici-
pated probability of exceedance of 10% for the capacity of the lateral load-resisting system
during the MCE in ASCE 7-16, Chapter 1. For TRC III structures, the importance factor of 1.25
results in a 5% exceedance probability given that the MCT has occurred, compared with a
corresponding anticipated exceedance probability of 5% to 6% for the lateral force–resisting
system during the MCE. Similarly, for TRC IV structures the probability of exceedance during

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Structural Design Procedures for Tsunami Effects 77

Table G8-2. Anticipated Reliabilities (Maximum Probability of Failure) for Earthquake and Tsunami Design.

Failure* probability conditioned on


Probability of failure* in 50-yrs Maximum Considered Event

Earthquake
Risk Category Earthquake Tsunami (MCE) Tsunami (MCT)
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

II 1% 0.3% 10% 7.5%


III 0.5% 0.2% 5% 5%
IV 0.3% 0.1% 2.5% 3%
Vertical evacuation N/A <0.1% N/A <1%
refuge structures
Source: Table 1.3-2; Chock et al. (2016).
*Note: Tsunami probabilities are based on exceeding an exterior structural component’s capacity that does not
necessarily lead to widespread progression of damage, but the seismic probabilities are for the more severe
occurrence of partial or total systemic collapse.

the MCT is 3% compared with 2.5% to 3% for the MCE. Finally, because of the increased con-
servatism in flow depth required for design of tsunami vertical evacuation refuge structures,
the exceedance probability is only 0.8% given that the MCT has occurred. This provides a
high degree of confidence that even during the most severe anticipated tsunami, those who
seek refuge in a building or other structure designed as a tsunami vertical evacuation refuge
following the provisions in ASCE 7-16, Chapter 6 will survive the tsunami.
Also shown in Table G8-2 are the probabilities of failure over a 50-year design life of the
building or structure. These probabilities are significantly lower for tsunami than seismic
conditions.

8.3.3  Load Combinations


Tsunami loads on the overall structural system and on individual structural elements must
be combined with other specified loads in accordance with the load combinations provided
in ASCE 7-16, Section 6.8.3.3 as follows:

0.9D + FTSU + HTSU (6.8-1a)

1.2D + FTSU + 0.5L + 0.2S + HTSU (6.8-1b)

where
D = Dead load,
L = Live load,
S = Snow load,
FTSU = Tsunami load effect for incoming and receding directions of flow, and

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


78 Tsunami Loads and Effects

HTSU = Load due to tsunami-induced lateral foundation pressures developed under


submerged conditions. Where the net effect of HTSU counteracts the principal load
effect, the load factor for HTSU shall be 0.9.

These load combinations are based on the extra­ordinary load combinations specified
in ASCE 7-16, Section 2.5 by replacing Ak with FTSU. Because tsunami design is based on the
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

maximum considered tsunami, the load factor is 1.0, as for other ultimate loading conditions
such as wind and seismic. Note that the tsunami loads are not combined with any seismic
ground shaking since the probability of a major aftershock occurring at the same time as the
peak tsunami wave loading is relatively remote, although a near-source tsunami will occur
some 15 to 60 min after a major subduction zone earthquake.

8.3.4  Structural System Acceptance Criteria


For evaluation of the overall lateral force–resisting system at the collapse prevention level,
which will satisfy the minimum requirement for TRC II and III buildings, ASCE 7-16 allows a
direct comparison between the tsunami base shear and the inelastic seismic capacity of the
lateral force–resisting system for Seismic Design Category D, E, or F. The inelastic seismic
capacity that can be utilized for this comparison is defined as 75% of the horizontal seismic
load effect, Emh, including the system’s overstrength factor, W0, as defined in ASCE 7-16, Chap-
ter 12. In other words, for TRC II and III structural systems designed to Seismic Design Cate-
gory D, E, or F, the following check satisfies the life safety level tsunami design requirement:

VTSU < 0.75Emh = 0.75W0Eh (G8-1)

Figure G8-3 shows this acceptance criterion graphically. Structural lateral force–
resisting systems designed to Seismic Design Categories D, E, and F, are designed for a
reduced seismic base shear, Eh, and then provided with significant nonlinear ductility through
careful attention to detailing of members and connections. A push-over analysis of a building
with this seismic design will yield a maximum apparent lateral load capacity that exceeds the
design base shear by a significant margin. The ratio between the apparent strength and the
design base shear is referred to as the overstrength
factor, W0. This overstrength factor is typically used
to ensure that structural elements required to remain ASCE 7 provides a high degree
elastic during an earthquake will indeed remain of confidence that even during
elastic even as the ductile components experience the most severe anticipated
their maximum inelastic response. As such, the tsunami, those who seek refuge
in a building or other structure
overstrength factor has been biased slightly high to
designed as a tsunami vertical
ensure limited damage to critical elements required evacuation refuge following the
to remain elastic. It may therefore overestimate the provisions in ASCE 7-16, Chapter
actual strength of the ductile framing system. 6 will survive the tsunami.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Structural Design Procedures for Tsunami Effects 79

When developing the ASCE 7-16 tsunami design provisions, consideration was given to
the potential that for near-source tsunamis, the tsunamigenic earthquake would cause dam-
age to the structure, which would then reduce the structural capacity remaining to resist
the tsunami loads. Assuming that the seismic capacity has been degraded somewhat during
a near-source tsunamigenic earthquake and recognizing that the tsunami loads are sus-
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

tained as opposed to rapid cyclic seismic loads, the acceptance criterion for tsunami base
shear was set at 0.75W0Eh. If the tsunami base shear is less than this limit, then the seismic
lateral force–resisting system does not need to be evaluated for tsunami loads. If the tsu-
nami base shear exceeds the seismic design base shear, the structural system will undergo
nonlinear deformations, resulting in the likelihood of a residual lateral displacement after
the event as illustrated by DR in Figure G8-3. However, if the tsunami base shear exceeds
this limit, then the seismic base shear capacity, Eh, must be increased until this criterion is
met. In other words, the lateral force–resisting system must be redesigned for a seismic base
shear of Eh = VTSU/0.75W0, distributed up the height of the building according to the original
earthquake load distribution. The resulting member forces in all structural members below
the maximum inundation height must be combined with the component loads attributable
to hydrodynamic drag and debris impact and designed for the increased member loads. The
appropriate seismic detailing must also be provided so that these members have the neces-
sary ductility and overstrength to meet the tsunami demand. Structural members above the
maximum inundation height do not need to be modified, but any transition from one floor to
the next must be evaluated for possible seismic vertical irregularity.
If the seismic design is for Seismic Design Category A, B, or C, then the lateral force–
resisting system must be explicitly evaluated and designed for the applied tsunami loads.
Because of the lack of special seismic detailing, the tsunami design cannot rely on significant
ductility beyond the seismic design capacity, Eh.

Figure G8-3. Schematic showing tsunami lateral force–resisting system acceptance criterion for TRC II
and III structures designed to Seismic Design Category D, E, or F.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


80 Tsunami Loads and Effects

For TRC IV structures, including vertical evacuation refuge structure, and TRC III Criti-
cal facilities, the preceding acceptance criterion is not applicable. For these cases the design
requirement is for immediate occupancy, so the lateral force–resisting system must be explic-
itly analyzed and the individual components of the system designed to resist the tsunami loads.
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

8.3.5  Structural Component Acceptance Criteria


In addition to verifying that the lateral force–resisting system (LFRS) is adequate for the
overall hydrodynamic drag on the structure, it is also necessary to evaluate all individual
members of the structural system, regardless of whether they are part of the LFRS or not.
The following three approaches are permitted for evaluating individual structural compo-
nents for tsunami loads and effects.

8.3.5.1  Acceptability Criteria by Component Design Strength


The first and simplest approach to evaluating an individual structural component for tsu-
nami loads and effects is to ensure that the design strength (which includes the applicable
material resistance factors, φ) is greater than the internal forces determined using a linearly
elastic, static analysis of the maximum considered tsunami loads and effects, considering the
appropriate load combinations of ASCE-17, Section 6.8.3.3.

8.3.5.2  Alternative Performance-Based Criteria


As an alternative to the linearly elastic, static analysis, one can also perform a nonlinear
static analysis using either an equivalent effective stiffness consistent with the secant
value at or near the yield point, or using a mathematical model incorporating the nonlin-
ear load-deformation characteristics of individual components of the structure, similar to
a push-over seismic analysis following ASCE 41 (2017b) provisions (Baiguero et al. 2020).
This performance-based design approach will generally provide a more economical design
than the prior linearly elastic static analysis approach because of the use of expected mate-
rial properties as defined in ASCE 41, instead of the nominal properties, and because this
approach utilizes the full nonlinear response of the structural members. For high-seismic
design regions, this nonlinear capacity can far exceed the elastic design capacity.
Nonlinear static lateral analysis, commonly known as push-over analysis, was developed
for seismic assessment of existing buildings, but can be used for design of new structures. It
involves the application of an incrementally increasing lateral load applied at each floor level
in accordance with the lateral load distribution defined by ASCE 7-16. The nonlinear prop-
erties of the structural materials are defined by ASCE 41. ASCE 41 also provides backbone
curves against which each structural element can be evaluated to determine its effective
strength and stiffness based on the displacement or rotation at each step of the push-over

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Structural Design Procedures for Tsunami Effects 81

analysis. These backbone curves are based on the hysteretic seismic response of experimen-
tal specimens representing the various yielding components of the structure. The actual
monotonic performance of a beam-column connection, for example, is generally slightly
stiffer and stronger than the backbone curve for the same specimen tested cyclically. Hence
the backbone curve is a conservative assumption for use in tsunami assessment, which is
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

essentially a statically applied increasing lateral load.


During tsunami inundation and drawdown, the structure must be assessed for varying
combinations of flow depth and velocity based on the actual time histories if a site-specific
tsunami hazard analysis has been performed, or using the normalized tsunami flow time
histories shown in Figure G8-2. If the nonlinear structural model is able to resist the lat-
eral loads resulting from the time history without experiencing a loss of lateral capacity
or developing a collapse mechanism, then the structural system is adequate to resist the
applied tsunami loads. However, all elements of the lateral force–resisting system must also
resist the component loads caused by hydrodynamic drag or debris impact loading. The
combined forces in the structural elements must be checked against the element capacity.
This includes checking the shear capacity of all structural elements because these are gen-
erally not automatically evaluated during a push-over analysis. Note that push-over analysis
must be performed for both incoming flow and during drawdown.
If this performance-based design approach is used for TRC IV buildings and structures,
then an independent peer review is required to ensure that the analysis requirements have
been applied correctly.

8.3.5.3  Alternative Acceptability by Progressive Collapse Avoidance


If the tsunami loads and effects exceed the capacity of a structural element, one can assume
that the element has failed and check the residual load-carrying capacity of the remain-
ing structure in accordance with the alternate load path procedure given in UFC 4-023-03
(DOD 2016). Progressive collapse-preventative design is also permissible when the design is
required to consider extraordinary impact loads according to Section 6.11.7.

8.4  Minimum Fluid Density for Tsunami Loads


Virtually all tsunami experimentation in laboratory settings utilizes clean fresh water. Except
for landslide-induced tsunamis in freshwater lakes, all tsunami flow consists of seawater,
which has a slightly higher density than fresh water. In addition, soon after reaching the
shoreline, the tsunami flow can pick up considerable amounts of sediment and smaller debris
items that increase its effective density. Therefore, for all hydrostatic and hydrodynamic
loading conditions in ASCE  7-16, Chapter 6, it is required that an increased fluid density
be used. At a minimum, this increase is achieved by applying a factor ks = 1.1 to the den-
sity of seawater (Review G8-1). This represents an equivalent sediment concentration of 7%,
assuming a specific gravity of 2.5 for the suspended soil particles as shown in Example G8-1.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


82 Tsunami Loads and Effects

Review G8-1. Fluid Density.

It is very important to correctly distinguish between fluid weight density, γ, and flu-
id mass density, ρ. Both values are used throughout ASCE 7-16, Chapter 6 and in this
guide.
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

The fluid mass density of seawater is

rsw = 2.0 slugs per cubic foot (sl/ft3) in US customary units


and
rsw = 1,025 kg/m3 in SI units.

Multiplying the mass density by g = 32 ft/s2 (9.8 m/s2), the acceleration due to
gravity, gives the fluid weight density of seawater as

gsw = 2.0 × 32 = 64 lb/ft3 in US customary units


and
gsw = 1,025 × 9.8 = 10.0 kN/m3 in SI units.

Example G8-1. Fluid Density Example 1.

Assuming a concentration of 7% sediment with a specific gravity of 2.5, the effective


fluid density factor is

ks = 0.93 × 1.0 + 0.07 × 2.5 = 1.1

If conditions at the site warrant consideration of a higher concentration of sed-


iment, then ks can be increased. For example, when considering a 10% sediment con-
centration, the effective fluid density factor would be

ks = 0.9 × 1.0 + 0.1 × 2.5 = 1.15

8.5  Flow Velocity Amplification


Tsunami inundation of an undeveloped coastline will generally take the form of an advanc-
ing surge, or possibly a bore, traveling at relatively uniform velocity across the advancing
front. This condition was observed during many of the aerial videos capturing the advance
of the tsunami over the Sendai coastal plain in Japan. However, when the flow interacts with

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Structural Design Procedures for Tsunami Effects 83

relatively sturdy buildings and other structures such as in an urban environment, there can
be considerable concentration of the flow leading to local velocity amplification.
This velocity increase can be considered by any of three approaches provided in
ASCE 7-16, Section 6.8.5 as follows:
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

1. Performing a site-specific inundation analysis in accordance with ASCE  7-16, Sec-


tion 6.7.6.6 but including modeling of the built environment. This approach requires a
high-resolution topographical model with a minimum resolution of 10 ft (3 m) so as to
accurately capture the buildings and other structures that may lead to flow amplification.
2. Using a Manning’s roughness factor of 0.04 for “buildings of at least urban density” from
ASCE  7-16, Table 6.6-1, when performing the energy grade line analysis according to
ASCE 7-16, Section 6.6.3.
3. Performing site-specific physical or numerical modeling in accordance with ASCE 7-16,
Section 6.8.5.2 or Section 6.8.10 as applicable.

The simplest of these options is clearly the selection of Manning’s n = 0.04 when per-
forming EGLA. This provides conservatism in both flow depth and velocity anticipated to
generally accommodate the effects of flow focusing.

8.5.1  Upstream Obstructing Structures


ASCE 7-16 requires that flow velocity amplification be considered, using one of the preced-
ing three methods, whenever enclosed structures of concrete, masonry, or structural steel
construction are located within 500 ft upstream of the site under consideration, and both of
the following apply:

1. These structures have a plan width greater than 100 ft or 50% of the width of the struc-
ture being designed, whichever is greater; and
2. The structures are located within an arc between 10 and 55 degrees either side of the
flow vector aligned with the center third of the width of the structure being designed.

These conditions are illustrated in ASCE 7-16 Commentary, Figure C6.8-3. Typical exam-
ples are shown in Figure G8-4 and Table G8-3. ASCE 7-16 Commentary, Section C6.8.5 also
provides useful results from a number of experimental studies on flow amplification caused
by the presence of rigid objects upstream of the structure under consideration. Consider-
ation should be given to potential future construction that could result in increased flow
amplification at the project site, but without prior knowledge of planned construction it may
not be possible to anticipate such effects.
Open structures conforming to the ASCE 7-16 definition in Section 6.2 need not be con-
sidered as obstructing structures.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


84 Tsunami Loads and Effects
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

Figure G8-4. Examples of upstream obstructing structures.


Note: See Table G8-3 for classification of upstream structures.

Table G8-3. Classification of Upstream Obstructing Structures in Figure G8-4.

Structure Amplifies
Upstream width flow
structure (ft) velocity Rationale
Width greater than 100 ft and 1/2 × 255 = 127 ft.
A 180 Yes Substantially within a 10 to 55 degree arc and within 500 ft
upstream.
Width greater than 100 ft and 1/2 × 255 = 127 ft. Fully within
B 135 Yes
a 10 to 55 degree arc and partially within 500 ft upstream.
Width greater than 100 ft and 1/2 × 255 = 127 ft.
C 180 Yes Substantially within a 10 to 55 degree arc and within 500 ft
upstream.
Width greater than 100 ft and 1/2 × 255 = 127 ft, but not
D 180 No
within a 10 to 55 degree arc.
E 70 No Fully within a 10 to 55 degree arc, but width less than 100 ft.
Fully within a 10 to 55 degree arc, but width less than 1/2 ×
F 110 No
255 = 127 ft.
Width greater than 100 ft and 1/2 × 255 = 127 ft, but outside
G 180 No
of a 10 to 55 degree arc.
Width greater than 100 ft and 1/2 × 255 = 127 ft.
H 135 Yes Substantially within a 10 to 55 degree arc and within 500 ft
upstream.
Width greater than 100 ft and 1/2 × 255 = 127 ft and within a
I 180 No
10 to 55 degree arc, but beyond 500 ft upstream.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Structural Design Procedures for Tsunami Effects 85

8.5.2  Flow Velocity Amplification by Physical


or Numerical Modeling
For more detailed analysis of the flow velocity amplification effects, it is also permissible to perform
site-specific numerical modeling as described in ASCE 7-16, Section 6.7.6, but with all obstructing
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

structures explicitly modeled. Alternatively, physical testing of the effects of obstructing struc-
tures can be performed in accordance with ASCE 7-16, Section 6.8.10. If this numerical model-
ing or physical testing indicates that the obstructing structures will reduce the flow velocity,
then their effect must be ignored. However, structural countermeasures specifically designed in
accordance with ASCE 7-16, Section 6.13 can be used to reduce the flow velocities.

8.6  Directionality of Flow


8.6.1  Flow Direction
Video evidence from past tsunamis has indicated that drawdown flow velocities may be as
large as, or even exceed, the incoming velocities, depending on the topographical conditions.
It is therefore important that both incoming and outgoing flow conditions be considered in
tsunami design. The energy grade line analysis is used to compute the estimated maximum
incoming flow velocity, which is assumed to be the same for the outgoing flow.
The energy grade line analysis relies on a 2D transect from the shoreline to the runup
point passing through the center of the project footprint. This central transect is oriented
perpendicular to a 1,000 ft length of shoreline, created by averaging the shoreline extending
500 ft on either side of the site (Figure G6-1). Because of 3D irregularities in the bathymetry
and topography around the site location, it is possible that the principal flow direction will
not be perpendicular to the shoreline. ASCE 7-16 therefore requires that all flow directions
within an arc measuring 22.5 degrees either side of the central transect be considered when
determining the flow conditions using EGLA. The transect within this arc having the largest
runup elevation may be used in EGLA to determine the flow depth and velocity for design
purposes at the site, which are then applied as both incoming and outgoing flow directions
within the ±22.5 degree arc. Alternatively, one can consider all possible directions within this
arc and perform EGLA for each critical direction, applying the resulting loads in that direc-
tion. This may require a number of individual EGLA and load determinations to ensure that
the worst case has been identified. Figure G6-1 illustrates this procedure.

8.6.2  Site-Specific Directionality


Site-specific tsunami inundation analysis is expected to provide flow directionality that
incorporates 3D effects of the bathymetry and topography affecting the project site. How-
ever, there is still some uncertainty in the accuracy of the model results, so a minimum vari-
ability of flow direction of ±10 degrees must be considered when applying the resulting loads
to the structure being designed.
@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


86 Tsunami Loads and Effects

8.7  Minimum Closure Ratio for Load Determination


The closure ratio, Ccx, represents the percentage flow obstruction presented by the building
structural elements and debris that accumulates on the exterior of the building (Figure G8-5)
(Chock et al. 2013b). Computation of Ccx is covered in ASCE 7-16, Section 6.10.2. Because of the
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

potential for significant debris accumulation along the perimeter of the structure, for both
incoming and outgoing flow directions, it is required that the minimum value of Ccx is 0.70,
unless the building qualifies as an open structure, in which case the minimum Ccx value is 0.50.
Example G8-2 shows how to determine whether a structure qualifies as an open structure.
The closure coefficient is given by ASCE 7-16 as

Ccx 
 A col  Awall   1.5 Abeam
(6.10-3)
Bhsx
where
Acol = Projected area of all structural columns below the flow depth,
Awall = Projected area of all structural walls below the flow depth,
Abeam = Projected area of the deepest submerged beam at each floor level, including
floor slab thickness, perpendicular to the flow direction,
B = Overall width of the building perpendicular to the flow, and
hsx = Height of the story being considered in the computation of Ccx.

Examples G8-3 and G8-4 demonstrate the calculation of the closure coefficient for the
prototype reinforced concrete residential building introduced in Chapter 2 when subjected
to flow in both the transverse and longitudinal direction. Note that Ccx can exceed 1.0 if there
are a large number of structural columns and/or walls within the building (Example G8-4) in
which case ASCE 7-16, Section 6.10.2.1 allows the use of Ccx = 1.0.

Figure G8-5. Debris damming on buildings in Otsuchi, Japan: (left) N 39.35968; E 141.90609; (right)
N 39.35783; E 141.90639.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Structural Design Procedures for Tsunami Effects 87

Example G8-2. Open Structure Example.

An open structure is defined as one in which the portion below the inundation depth
has no greater than 20% closure ratio and also has no tsunami breakaway walls or inte-
rior partitions. Open structures need not consider LC1 buoyancy or scour effects and
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

have a reduced potential for debris accumulation (minimum Ccx = 0.5 instead of 0.7 for
regular buildings). Examples of open structures would be one with parking levels at the
base of a building, or a hotel with an open-air atrium.
Whether or not a structure qualifies as an open structure depends on the ratio
of the vertical projected area of structural elements to the total inundated vertical
projected area. This example shows the calculation of the projected area ratio for the
reinforced concrete office and residential buildings introduced in Chapter 2, assuming
a flow depth of 25 ft and a flow direction perpendicular to the broad side of the build-
ing. Refer to Figure G2-3 and Figure G2-4 for dimensions of the structural frames.
For the moment-resisting frame office building

 30in. 24in. 
Structural projected area = 10 × + 2× × 255ft = 725ft
2
 12 12 

and total projected area = 254 ft × 25 ft = 6,350 ft2

725
Therefore, the projected area ratio = = 0.114 < 0.20
6, 350
This would qualify as an open structure if the lower two levels had no exterior
walls or interior partitions.
For the shear wall residential building

 20in. 
Structural projected area = 6 × + 2 ×10ft + 28ft × 25ft = 1, 450ft 2
 12 

and total projected area = 254 × 25 ft = 6,350 ft2

1, 450
Therefore, the projected area ratio = = 0.228 > 0.20
6, 350
This would not qualify as an open structure even if the lower three levels had no
exterior walls or interior partitions.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


88 Tsunami Loads and Effects

Example G8-3. Closure Coefficient Example 1.

The closure coefficient, Ccx, is required for calculations of drag force on the overall
structure and individual exterior structural elements. This example shows the cal-
culation of Ccx for the overall structural load assessment of the reinforced concrete
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

residential building introduced in Chapter 2, assuming a flow depth of 25 ft and a flow


direction perpendicular to the broad side of the building. Refer to Figure G2-4 for
dimensions of the structural frame.
ASCE 7-16 gives

Ccx 
 A col  Awall   1.5 Abeam
(6.10-3)
Bhsx

To compute the Ccx value for lateral load at the second level, the vertical projected
area of all structural elements between midheight of the floor below and floor above
must be included. This represents a height hsx = (12 + 9 ft)/2 = 10.5 ft. Therefore,

 20in.  8in.  
Acol  32    10.5ft     524ft
2

 12  12 

  8in.  
Awall  2   (10  28ft )  10.5ft     747ft
2

  12 

8in.
Abeam   254  169ft 2   (for slab edge)
12
B = 254 ft

and hsx = 10.5 ft

Therefore, Ccx 
  524  747   1.5 169  0.57  0.70
254 10.5
So Ccx = 0.70

Because this building has exterior walls and interior partitions, it does not qualify
as an open structure (see Example G8-2). Therefore, the minimum Ccx = 0.70 must be
used for computing lateral loads on this floor level.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Structural Design Procedures for Tsunami Effects 89

Example G8-4. Closure Coefficient Example 2.

This example shows the calculation of Ccx for the overall structural load assessment of
the residential building introduced in Chapter 2, assuming a flow depth of 25 ft and a
flow direction perpendicular to the narrow side of the building. Refer to Figure G2-4
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

for dimensions of the structural frame.


ASCE 7-16 gives

Ccx 
 A col  Awall   1.5 Abeam
(6.10-3)
Bhsx

To compute the Ccx value for lateral load at the second level, the vertical projected
area of all structural elements between midheight of the floor below and floor above
must be included. This represents a height hsx = (12 + 9 ft)/2 = 10.5 ft. Therefore,

 20in.  8in. 


Acol = 32 × ×10.5ft −  = 524 ft
2
 12  12 

  8in. 
Awall = 2 ×(16 + 11ft ) ×10.5ft −  = 531 ft
2


  12 

8in.
Abeam = × 64ft = 42.7ft 2   (for slab edge)
12

B = 64 ft

and hsx = 10.5 ft

Therefore, Ccx =
∑ (524 + 531) + 1.5× 42.7 = 1.67 > 1.0
64 ×10.5
Therefore, according to ASCE 7-16, Section 6.10.2.1, Ccx = 1.0 may be used for com-
puting lateral loads on this floor level.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


90 Tsunami Loads and Effects

8.8  Minimum Number of Tsunami Flow Cycles


Tsunami inundation typically consists of a series of waves, in which the first may not be the
largest. To incorporate the deleterious effects of multiple waves, ASCE 7-16 requires that at
least two tsunami flow cycles be considered. The first is based on 80% of the MCT, and the
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

second represents the MCT inundation. The primary effect of the initial 80% MCT inunda-
tion will be to induce scour around the building foundations. Lateral loads applied by the
subsequent MCT inundation must now be resisted by the foundation system assuming that
scour has already occurred. It is unlikely that this repeat loading will affect any of the struc-
tural member designs, but it may impact the design of the foundation elements.

8.9  Seismic Effects on the Foundations


Preceding Local Subduction Zone Maximum
Considered Tsunami
For near-source tsunamis, in which the tsunami is the result of an offshore subduction zone
earthquake, all structures in the coastal region will have experienced significant ground
shaking during the earthquake. In particular, for RC II and III buildings the occurrence of the
MCE would represent ground shaking exceeding the design level. The effect on the structure
itself is incorporated into the lateral load evaluation in ASCE 7-16, Section 6.8.3.4 by allowing
only 75% of the structural seismic lateral load capacity, including overstrength, to be uti-
lized for tsunami lateral loads provided the structure was designed and detailed for Seismic
Design Category D, E, or F.
Ideally, immediately after the earthquake shaking stops, all those in the tsunami evac-
uation zone will move to higher ground or to designated tsunami vertical evacuation refuge
structures. ASCE 7-16 requires that these structures be designed as Seismic Risk Category
IV and Tsunami Risk Category IV. As a result, these structures are expected to be available
for immediate occupancy after the earthquake. However, many evacuees may not respond
immediately to the ground shaking or tsunami warnings and may seek refuge in taller Risk
Category II or III buildings in the inundation zone. These buildings will have experienced
significantly more seismic damage, especially if the earthquake represents the MCE. Risk
Category II buildings can be expected to be at the collapse prevention performance level,
whereas Risk Category III buildings would be at the damage control performance level.
Hopefully evacuees who are not able to reach high ground or a designated refuge structure
would self-select a taller building with less visible seismic damage as a refuge of last resort.
However, the large-magnitude earthquake can also have significant effects on the build-
ing foundations owing to liquefaction, lateral spreading, and other geotechnical effects com-
mon to large earthquakes. It is required that the geotechnical investigation report evaluate
seismic effects on the foundation such as slope instability, liquefaction, total and differential
settlement, surface displacement caused by faulting, and lateral spreading and lateral flow.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Structural Design Procedures for Tsunami Effects 91

Any anticipated geotechnical effects of the earthquake must be considered as the starting
point for analysis of tsunami-induced pore pressure softening, scour, and so forth.

8.10  Physical Modeling of Tsunami Flow,


Loads, and Effects
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

Physical modeling is permitted as an alternative to the prescriptive procedures provided


for flow amplification (ASCE 7-16, Section 6.8.5), hydrodynamic loads (Section 6.10), debris
impact loads (Section 6.11), and foundation design (Section 6.12). This section provides the
requirements that govern this physical modeling to ensure that the results are reliable for
use as replacements for the prescriptive procedures. It is particularly important that the
minimum scale factors provided in ASCE 7-16, Table 6.8.3 are satisfied to ensure that scale
effects do not distort the test results. Debris impact testing also must be performed at full
scale to ensure accurate impact load and duration results.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

Tsunami Loads and Effects


@Seismicisolation
@Seismicisolation
9
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

Hydrostatic Loads

Hydrostatic loads to be considered in tsunami design include buoyancy, unbalanced lateral


hydrostatic force, residual water surcharge on floors and walls, and hydrostatic surcharge
pressure on foundations. The various hydrostatic loading expressions are described in this
chapter and demonstrated using examples.

9.1  Buoyancy
All objects submerged in water experience an uplift force equal to the weight of water dis-
placed by the object. For buildings subjected to tsunami inundation this is particularly crit-
ical when structural and/or nonstructural walls prevent water from entering an enclosed
space in the building.
A number of concrete and structural steel framed buildings were observed to float
during the Tohoku tsunami because of buoyancy effects. Survivor videos show a number
of virtually intact multistory buildings in Onagawa, Japan, floating in approximately 20 ft
(6 m) of water during the Tohoku tsunami. Some of these buildings may be timber framed,
but a number were structural steel and concrete buildings. Chock et al. (2013b) analyzed a
reinforced concrete building used for refrigerated storage of fish products (Figures G8-1
and G9-1). Because the lower level of the building housed a large sealed refrigerator storage
room, uplift caused by buoyancy was sufficient to float the building, remove it from its com-
pression-only piles, and transport it about 50 ft (15 m) inland where it flipped over and came
to rest on its side. Based on measurements of the building, it was predicted to float when the
water level reached 23 ft (7 m). In fact, the water level reached over 62 ft (19 m), easily over-
topping the 2-story building (Chock et al. 2013b).
A building will only experience these overall uplift forces if the water has time to perme-
ate through the soil below the building and elevate the sub-slab pressure to match the adja-
cent hydrostatic pressure. Figure G9-2 shows two potential pressure distributions below
the ground-level slab. For a permeable soil, the conservative assumption is that hydrostatic
pressure develops below the foundation slab as shown by the dashed line. Less permeable
soil will delay the pressure development resulting in a pressure profile similar to the solid

93
@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


94 Tsunami Loads and Effects
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

Figure G9-1. Reinforced concrete fish storage building in Onagawa, Japan, that floated off its
compression pile foundations during the Tohoku tsunami.

line. Because the typical period of tsunami waves is on the order of 10 to 30 min, it is rea-
sonable to assume that foundation soils such as silts, sands, and gravels will allow significant
pressure to develop below the ground-level slab. More cohesive soils such as clay and clayey
silt will require longer inundation times for the pressure to equalize. However, if a gravel
or sand layer is placed below the ground-level slab, which is common practice, this layer
will be permeable and could allow pressure to develop more quickly below the slab. In the
case of the Onagawa, Japan, refrigerator building, coarse gravel below the foundation slab
would have provided ample permeability for pressures to equalize almost immediately (Fig-
ure G8-1).
Note that in the Onagawa, Japan, building shown in Figure G8-1, the slab-on-grade was
integrally connected to the grade beams, therefore pressure that developed below the slab
applied uplift to the entire foundation system and building above. If the ground-level slab
was a lightly reinforced slab-on-grade with isolation joints at all columns and grade beams,
then the pressure below the slab would lift the slab only and not the building. This would
allow water to enter the ground floor and reduce the buoyancy effect. In many cases this
will be the simplest option for relieving overall uplift caused by buoyancy (Figure G9-3).

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Hydrostatic Loads 95
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

Figure G9-2. Potential hydrostatic pressure corresponding to Load Case 1 buoyancy calculation.

Figure G9-3. Isolated ground floor slab-on-grade to reduce buoyancy effect.

Resulting damage to the slab-on-grade does not jeopardize the structural integrity of the
building, although it may become tumbling debris that could strike ground floor columns
as witnessed in Samoa (Figure G9-4). The impact caused by tumbling debris is addressed in
ASCE 7-16, Section 6.11.4.
Nonstructural exterior walls that are specifically designed to break away under tsu-
nami loads, as well as standard glazed windows, can be assumed to fail once the external
hydrostatic pressure exceeds their capacity and allow water into the building interior, thus
relieving the buoyancy effect. However, standard exterior walls consisting of metal or wood
studs, or concrete masonry units (CMUs), will not break away at a predictable load unless
specifically designed to do so. Continuity of the wall framing past the floor slabs results in
considerable lateral capacity that is generally not considered when the exterior walls are
designed for wind load. Kleinman et al. (2007) tested standard metal stud wall framing with
conventional track at the base and slip track at the top of the wall. The lateral load required

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


96 Tsunami Loads and Effects
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

Figure G9-4. Tumbling concrete debris from slab-on-grade at a building in Malaela, Samoa (S 14.0203;
W 171.4250).

to fail these systems was up to 3.5 times greater than their design capacity. Designing an
exterior wall to resist 30 psf wind load could therefore result in a wall with ultimate capacity
of over 100 psf, which would have to be exceeded for it to break away. Kleinman et al. (2007)
proposed installing a simple “fuse” at the top of each stud to ensure failure at a predictable
lateral load.
Nonstructural CMU infill walls are often used at the ground-floor level for security or
sound abatement. These walls are generally reinforced with both vertical and horizontal
steel reinforcing bars. They are typically isolated from the columns and beam above so as not
to influence the seismic performance of the steel or concrete frame, but they are commonly
braced laterally at the second-floor beams to resist out-of-plane loads caused by wind and
seismic effects. For CMU walls to break away under tsunami loads, it might be necessary to
remove the top lateral bracing and isolate the wall footing so that it is able to rotate and allow
the wall to collapse at a predictable lateral load. The wall would still need to be stable during
design wind and seismic events, so designing a breakaway CMU wall may be more difficult
than for stud-framed walls.
Breakaway walls must therefore be designed to resist the required wind and seismic
loads but have limited reserve capacity prior to failure. To accommodate potential over-
strength of the nonstructural walls, ASCE 7-16 requires that designers consider a minimum
uplift from buoyancy, along with the associated hydrodynamic drag, as in Load Case 1 (see

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Hydrostatic Loads 97

Section 8.3.1 of this guide). One is allowed to assume that exterior windows with standard
glazing will fail when the water level reaches the top of the window. Alternatively, particu-
larly for tall windows, one is permitted to compute the water level at which the unbalanced
lateral hydrostatic pressure will be sufficient to break the glazing. Example G9-1 shows a
comparison between wind and hydrostatic load for a typical ground-floor window. However,
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

windows designed for large missile wind-borne debris impact, such as in hurricane zones,
and those designed for blast loading will have significantly larger pressure capacity than
standard glazing and should be assumed to remain intact throughout the tsunami.
Even with a breakaway ground-floor slab or wall system, buildings will still experience
buoyancy caused by water displacement by the structural framing members, combined with
hydrodynamic lateral loads. The failure of a 2-story reinforced concrete framed building
in Otsuchi, Japan, is attributed to this combination of effects, even though the buoyancy
was significantly reduced by the failure of the timber framing at the ground-floor level
(Figure G9-5).
The buoyancy effect must include the potential for air trapped below the floor slab by
beams and joists, and floor areas that have not yet been inundated. Figure G9-6 shows an
example in which the water level has risen above the second floor of a reinforced concrete
building. Water is free to enter the ground floor because of failure of the nonstructural walls
and windows, or uplift of the nonstructural slab-on-grade. However, air will become trapped
below the second floor by the presence of beams and joists supporting the floor. This will
occur in both reinforced concrete and structural steel buildings with beams supporting the
floor slabs but would not occur for a flat slab system without beams. The highest beam soffit
will define the bottom elevation of the trapped air in any particular floor panel. In reality,
some of this air will be driven out by turbulence in the water. Hydrostatic pressure will also

Figure G9-5. Two-story building in Otsuchi, Japan, that overturned because of combined buoyancy and
hydrodynamic lateral loads, even though the ground-floor timber framing was not able to resist the
hydrostatic uplift (N 39.35894; E 141.91128).

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


98 Tsunami Loads and Effects

Example G9-1. Window Pressure Comparison.

The figure shows a typical ground-floor window assumed to be 4 ft (1.2 m) high by 8 ft
(2.4 m) wide. For simplicity, the glass is assumed to span vertically to resist lateral load.
A typical coastal wind pressure on components and cladding, such as a window, would
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

be on the order of 38 psf (1.8 kN/m2). If the water level has reached the top of the win-
dow, then the hydrostatic pressure distribution on the glazing will be triangular, vary-
ing from zero at the top of the window to a maximum at the bottom.
The pressure at the bottom of the window is given by

p = gsbh = (1.1 × 64) × 1 × 4 = 282 psf

1.2
p   s bh  1.11, 025  9.81 1  13.3 kN / m 2
1, 000

The maximum bending moment in the glass from the assumed wind pressure is

pl 2 42  1.22 
M max   38   76ft-lb  1 .8   0.324 kNm 

8 8  8 

The maximum bending moment in the glass owing to the hydrostatic pressure is

1 
M max  0.128Wl  0.128    282  4   4  289ft-lb 1.226 kNm 
2 

With the water pressure inducing a bending moment almost four times that of the
wind load, along with the presence of floating debris in the water, it is reasonable to
assume that the glazing will fail once the water level exceeds the top of the window.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Hydrostatic Loads 99
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

Figure G9-6. Example of buoyancy caused by trapped air and floors that have not been inundated.

cause a slight reduction in the trapped air volume owing to compressibility. However, these
effects can be ignored to obtain a slightly conservative estimate of the displaced volume of
water.
As the water level rises above the second floor, it will not automatically inundate the
interior volume since the exterior wall system will keep the water out until it fails. This fail-
ure can be estimated based on the design wind load for which the exterior wall and any
windows were designed. Example G9-1 shows a typical comparison between wind load and
water load on an exterior window. ASCE 7-16 provides no guidance on the realistic estima-
tion of breakaway strength of nonstructural exterior walls, which can often be substantially
greater than the required strength for wind. Tests on single-span metal stud–framed walls
showed that the actual failure load could be three to four times the required design strength
because of conservatism in the design approach, material overstrength, and redundancies in
the system that are ignored for simplicity during the design process (Kleinman et al. 2007).
If the studs are continuous past the edge of the floor slab but are designed for wind load as
simply supported between floors, this will add significantly to the effective overstrength. A
procedure similar to Example G9-1 could be used to determine the maximum water level at
which the second floor exterior wall system will fail, thereby defining the displaced volume
for the buoyancy check shown in Figure G9-6. Note that the volume of all structural elements
below the water level, including grade beams, must be included in the buoyancy calculation.
Once the governing water depth has been determined for each load condition, the total
uplift force applied to the building, Fv, caused by buoyancy is given by

Fv = gsVw (6.9-1)

where gs = ksgsw is the weight density of seawater including sediment effect, and Vw is the total
volume of water displaced by the building.
Example G9-2 shows application of the buoyancy calculation to a submerged second
floor slab, and Example G9-3 shows application of Load Case 1 to a reinforced concrete
frame building with a 1-story basement.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


100 Tsunami Loads and Effects

Example G9-2. Exterior Nonstructural Wall Failure Load.

Consider the 6-story office building shown in Figure G2-3 and determine the water
depth for which the nonstructural exterior walls at the second level will fail.
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

Assume that the exterior walls are framed using metal stud walls spanning 9.83 ft
(3 m) between floor slab and the beam above and that they are designed to resist 38 psf
(1.8 kN/m2) wind pressure. Then the water height required to fail these walls (assuming
an overstrength factor of 3.0) would be determined as follows:

Wind load per foot width of exterior wall is Fw = 38 × 9.83 = 373 plf (= 1.8 × 3 = 5.4 kN/m)

Shear at bottom of wall owing to wind load is Vw = 373/2 = 186 plf (= 5.4/2 = 2.7 kN/m)

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Hydrostatic Loads 101

The hydrostatic pressure at the bottom of the window is given by

p = gsbh = (1.1 × 64) × 1 × h = 70.4h psf

[p = gsbh = (1.1 × 1,025 × 9.81) × 1 × h/1,000 = 11.1h kN/m2]


Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

Treating the nonstructural wall as simply supported, the shear at the bottom of
the wall caused by hydrostatic pressure, Vh, can be determined by taking moments
about the top of the wall, giving

1  h
Vh  9.83  p  h   9.83    0 , where p = 70.4h psf
2  3

1 1  h
Therefore, Vh    70.4h  h   9.83    35.2h 2  1.20h3
9.83 2  3

To fail the bottom connection, Vh ≥ 3Vw, assuming an overstrength factor of 3.0.


Therefore, failure is predicted to occur when 35.2h2 – 1.20h3 = 3 × 186 = 558 psf
(26.7 kN/m2).
Solving this equation gives h = 4.31 ft (1.31 m) above the second-level slab. This will
result in a large volume of displaced water causing buoyancy of the second-floor slab.
Having determined the water level representing the maximum depth for which
the second floor will remain dry, one can determine the hydrostatic uplift on the sec-
ond-floor slab as p = gs(h+d) as illustrated in Figure G9-6. For example, with 26 in.
(0.66 m) deep beams and a 4.31 ft (1.31 m) height required to fail the exterior walls at the
second level, the uplift pressure on the soffit of the second-floor slab would be

 26 
p   s h  1.1 64     4.31  456 psf
 12 

 (0.66  1.31) 
 p   s h  1.11, 025  9.81  1, 000  21.8 kN/m 2 
 

Compared with a typical 8 in. (203 mm) concrete floor slab with a dead weight of
100 psf (4.8 kN/m2), it is clear that this uplift pressure far exceeds the self-weight of the
slab. If the second floor slab is critical for bracing the vertical load-bearing elements,
such as columns and walls, or for transferring tsunami lateral loads to the lateral force–
resisting system, then it will need to be designed for this uplift pressure.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


102 Tsunami Loads and Effects

Example G9-3. Load Case 1: Buoyancy Plus Lateral Hydrodynamic Load.

Figure G9-7 shows a typical reinforced concrete frame building with a 1-story base-
ment. Assuming that the soil surrounding the basement is permeable, or that drainage
material has been installed against the outside of the basement walls, and a sand layer
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

below the basement slab, hydrostatic pressure will develop on the side walls and base
slab of the basement (Figure G9-8). If the base slab is designed to be integral with the
grade beams, walls, and columns, then the basement will act as an enclosed volume
subject to uplift. Even if there are leaks into the basement, the rate of those leaks is
unlikely to match the rate of increase in the exterior water level during a tsunami.
Unless the soil below the foundation was already saturated before the tsunami, there
will be air trapped within the soil below the foundation slab and between the grade
beams. The volume of displaced water will therefore include the soil between grade
beams and below the basement slab (Figure G9-8).

Figure G9-7. Building with 1-story basement.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Hydrostatic Loads 103
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

Figure G9-8. Hydrostatic pressure on building with 1-story basement.

Load Case 1 requires consideration of the combined effects of buoyancy and lat-
eral hydrodynamic load when the exterior inundation depth reaches the lesser of one
story or the height of the top of the first-story windows, provided this does not exceed
the maximum inundation depth at the site. For Figure G9-8, the top of the window ele-
vation of 9.5 ft (2.9 m) above grade is the controlling inundation depth for Load Case 1.
The pressure at the bottom of the foundation grade beams is given by

p = gsh = (1.1 × 64) × (9.5 + 13.5) = 1619 psf

 (2.9  4.25) 
 p   s h  1.11, 025  9.81  1, 000  79.1 kN/m 2 
 

The exterior wall of the basement must be designed to resist this hydrostatic lat-
eral pressure in addition to any lateral earth pressure from the saturated soil.
If the soil below the building is porous, or a sand layer is installed below the base-
ment slab, the hydrostatic pressure will develop across the entire basement slab (Fig-
ure G9-8). The buoyancy can be computed from the displaced volume, Vw, or from

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


104 Tsunami Loads and Effects

the base pressure acting upward over the area of the basement slab. This only applies
because the building has a uniform plan cross section below the water level. The uplift
force can therefore be computed in either of the following ways:

1. Using Equation (6.9-1)


Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

Fv = gsVw = (1.1 × 64) × (13.5 + 9.5) × (71 × 71)/1,000 = 8,161 kips

[Fv = gsVw = (1.1 × 1,025 × 9.81) × (4.25 + 2.9) × (21.6 × 21.6)/1,000 = 36,900 kN]
2. Computing total uplift pressure on basement slab

Fv  pA  1619 
 71 71  8,161 kips
1, 000

 Fv  pA  79.1  21.6  21.6   36, 900 kN 


Because the self-weight acts to oppose the uplift force from buoyancy, the load
combination given by Equation (6.8-1a) applies

0.9D + FTSU + HTSU (6.8-1a)


where
D = Building dead weight,
FTSU = Buoyancy force, and
HTSU = 0, as there is no differential pressure across the bottom of the basement slab.

The total dead weight of the building is approximately 5,115 kips (22,750 kN), so
0.9D = 4,600 kips (20,475 kN), which is less than the uplift force. Therefore, the piles
must be designed for a minimum tensile capacity of Tpiles = 8,161 – 4,600 = 3,561 kips
(Tpiles = 36,900 – 20,475 = 16,425 kN). This equates to 225 kips (1,030 kN) ultimate design
tension capacity for each of the 16 piles. It must be noted that the pile capacity may be
affected by pore pressure softening and local scour addressed in ASCE 7-16, Sections
6.12.2 and 6.12.4, respectively. If the tsunami is generated by a near-source earthquake,
then liquefaction effects on the pile tensile capacity must also be considered.
Simultaneous with the buoyancy force, Load Case 1 requires consideration of lat-
eral hydrodynamic load corresponding to the same inundation height of 9.5 ft (2.9 m)
above grade. For this example, the maximum inundation depth is taken as hmax = 28 ft
(8.5 m) and the maximum velocity is umax = 15 ft/s (4.6 m/s). The flow velocity corre-
sponding to h = 9.5 ft (2.9 m) can be determined from ASCE 7-16, Figure 6.8-1 as shown
in Figure G9-9. Entering the normalized inundation depth plot with h/hmax = 9.5/28
= 0.34 (h/hmax = 2.9/8.5 = 0.34), the corresponding flow velocity ratio is u/umax = 0.80.
Therefore, u = 0.8 × 15 = 12 ft/s (u = 0.8 × 4.6 = 3.68 m/s).
As shown later in this guide (Section 10.2.1), the hydrodynamic drag on the overall

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Hydrostatic Loads 105

building is given by ASCE 7-16 as

1
Fdx   s ITSU Cd Ccx B(hu 2 ) (6.10-2)
2
where
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

rs = 1.1 × 2.0 = 2.2 sl/ft3 (rs = 1.1 × 1,025 = 1127 kg/m3),


ITSU = 1.0 (Table 6.8-1: TRC II),
Cd = 1.4 (Table 6.10-1 based on B/hsx = 254/9.5 = 26.7),
Ccx = 1.0 since the exterior walls are assumed to be intact for Load Case 1,
B = 71 ft (21.6 m) overall width of building, and
h = 9.5 ft (2.9 m).

Therefore,

1
Fdx   2.2  1.0  1.4  1.0 

71 9.5  122 
 150 kips
2 1, 000

 1
 Fdx  1,127 1.0 1.4 1.0 

21.6 2.9  3.682
 669 kN 


 2 1, 000 

The lateral force–resisting system, including foundations, must be capable of


resisting both the net uplift and hydrodynamic lateral load acting simultaneously.

9.2  Unbalanced Lateral Hydrostatic Force


Differences in water level on opposite sides of a wall will result in unbalanced hydrostatic
forces (Figure G9-10 and Example G9-4). This will occur regardless of the wall orientation
relative to the tsunami flow. However, if water is able to pass around or through the wall so
as to limit the difference in water levels, this hydrostatic force will be reduced. ASCE 7-16
requires that unbalanced hydrostatic forces be considered for all structural walls that are
greater than 30 ft (9.1 m) wide without openings, or with openings that do not exceed 10%
of the wall area. This force must also be considered to act on walls that consist of two- or
three-sided perimeter structural walls, regardless of the wall length, because of the likeli-
hood that water pressure will develop on only one side of the wall. Figure G9-11 shows typical
wall configurations for which hydrostatic lateral loads must be considered. ASCE 7-16 only
requires this loading condition to be considered during Load Cases 1 and 2 for incoming flow
(see ASCE 7-16, Section 6.8.3.1 for load cases). Nevertheless, consideration should be given
to the potential for water to be trapped behind a structural wall during rapid drawdown as
shown in Figure G9-10.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


106 Tsunami Loads and Effects
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

Figure G9-9. Flow velocity determination for Example G9-3 Load Case 1 flow depth.
Source: ASCE 7-16, Figure 6.81.

Figure G9-10. Hydrostatic pressure that can develop on either side of a wide wall during incoming or
outgoing flow.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Hydrostatic Loads 107

Example G9-4. Unbalanced Hydrostatic Load.

A cantilever structural wall is subjected to a maximum flow depth of 9.8 ft (3 m) as


shown in Figure G9-10. The total hydrostatic force is given by ASCE 7-16 as
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

1 2 (6.9-2)
Fh = g s bhmax
2
where
gs = ksgsw = 1.1 × 64 = 70.4 lb/ft3 ( = 1.1 × 10.0 = 11.0 kN/m3),
b = 1 ft (1 m) for load per unit width of wall, and
hmax = 9.8 ft (= 3 m).

1
Therefore, Fh = × 70.4 ×1× 9.82 = 3, 380 lb / ft width
2
 1 
 Fh = ×11.0 ×1× 32 = 49.5 kN/m width 
 2 

This resultant force will act at 1/3 hmax = 3.3 ft [1 m] above grade. The wall stability
must be checked for overturning and sliding. It may be necessary to include a shear key
to generate adequate passive pressure to augment the base friction to resist sliding.
The soil properties also need to be adjusted for the effects of pore pressure softening
and potential scour of the soil over the footing (see Chapter 12 of this guide).

Figure G9-11. Wall configurations that must consider hydrostatic lateral load.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


108 Tsunami Loads and Effects

9.3  Residual Water Surcharge Load on Floors and


Perimeter Walls
During drawdown, water will drain from inundated buildings. If there are obstructions to this
drainage, water may become temporarily trapped on top of an elevated floor as the water below
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

the floor recedes. This will result in unbalanced vertical pressures on the top and bottom of the
floor slab equivalent to the weight of water retained on the slab. Given that the density of seawater
with sediment and debris is taken as 1.1 × 64 = 70.4 pcf (11 kN/m3), just 1 ft of retained water results
in a gravity load on the slab that exceeds the 40 or 50 psf (2 or 2.5 kN/m2) specified as minimum
distributed live loads for residential and office occupancies, respectively (ASCE 7-16, Table 4.3-1).
Water will be retained on an elevated slab if there are structural components such
as upturned beams that restrict water drainage from the slab. ASCE 7-16 requires that the
designer verify the slab can support a water depth equal to the height of any “continuous
perimeter structural components that have the capacity to survive the tsunami loads, and
that would retain water on an inundated floor.” It is assumed that all nonstructural walls,
windows, and other exterior finish and interior partitions will have been breached by the
incoming flow, otherwise there would not be any water on the slab under consideration.
However, even though the nonstructural wall elements have failed, any structural elements
that are expected to survive the tsunami flow will tend to retain water. There will likely be
some gaps in these structural elements at stairs and elevators, but this limited drainage
may not be sufficient to prevent water retention on the slab. Figure G9-12 shows damage
to the reinforced concrete beams supporting a second-floor slab in Taro, Japan, after the
Tohoku tsunami. This damage is attributed to the weight of water retained on the second
floor during drawdown, although subsequent waves have removed any perimeter wall that
may have caused the water retention.

Figure G9-12. Residual water on second-level floor slab is suspected of causing flexural cracks in
reinforced concrete floor beams, Taro, Japan (N 39.7374; E 141.9754). The original timber-framed second
floor and roof were destroyed during the tsunami but may have retained sufficient water to cause the
observed damage to floor beams.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Hydrostatic Loads 109

Example G9-5 shows a typical condition of residual water surcharge loading on an ele-
vated slab and the surrounding walls.

Example G9-5. Slab Surcharge Example.


Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

Consider the same multistory building used for the buoyancy check in Example G9-3.
Assuming that the maximum inundation depth at the site was 26 ft (7.93 m), the first
two floors would have been inundated. During drawdown, there is a tendency for
water to be retained on each floor level, particularly if it is surrounded by upturned
beams, structural parapets, or partial height walls. Assuming that CMU or concrete
walls or upturned beams have been used up to the windowsill level, then the potential
for retained water must be considered on both the first and second floors. If the first
floor were a slab-on-grade, this surcharge would not be an issue for the ground floor.
However, because the basement cannot be assumed to be fully flooded after draw-
down, it would be necessary to consider a surcharge load from the retained water on
the ground floor slab.
ASCE 7-16 gives the surcharge pressure as

pr = gshr (6.9-3)

where hr = hmax – hs, and hs = top of floor slab elevation; but hr need not exceed the
height of the continuous portion of any perimeter structural element at the floor.
From Figure G9-13, hr = 3 ft (0.9 m) for both floors, representing the height of
the concrete wall around the perimeter of each floor level. Even though there may be
some gaps in this wall for doorways, in particular at the ground-floor level, the rate of
drawdown during a tsunami is often rapid enough that all the retained water will not
be able to flow out of these openings in time to relieve the surcharge pressure. It is
therefore suggested that the conservative assumption be made that the water is fully
retained unless a specific flow analysis is performed to determine the residual water
on each floor level during drawdown.
Therefore, the surcharge pressure that develops on both ground- and second-floor
levels in Figure G9-13 is

pr = gshr = 1.1 × 64 × 3 = 211 psf

[pr = gshr = 1.1 × 1,025 × 9.81 × 0.9/1,000 = 9.95 kN/m2]

These floors must then be designed for combined dead load and this surcharge
load. Noting that most office and residential buildings are designed for between 40
and 100 psf (2 and 5 kN/m2), this retained water surcharge is considerably higher and
will likely govern the floor framing design. To avoid retaining water on elevated floors,
it would be advisable to avoid structural walls or upturned beams around the edge of
slabs at levels below the tsunami inundation height.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


110 Tsunami Loads and Effects
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

Figure G9-13. Water retained on elevated slabs during drawdown.

9.4  Hydrostatic Surcharge


Pressure on Foundations
During both incoming and outgoing tsunami flow, it is possible to develop differences in
water elevation on opposite sides of a building, wall, or other structural element that sur-
vives the flow loading. This difference in water elevation will result in differential surcharge
pressure on the foundation as shown schematically in Figure G9-14. Design of the founda-
tion must accommodate the potential for this differential surcharge.

Figure G9-14. Foundation surcharge resulting from unequal water depths on opposite sides of
structural wall.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


10
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

Hydrodynamic Loads

Hydrodynamic loads include drag forces on entire buildings and individual components,
impulsive lateral forces on walls, and uplift on elevated slabs and piers. This chapter intro-
duces the various hydrodynamic loading expressions along with a description of the lab-
oratory experiments and field observations that led to their development. Each loading
expression is demonstrated through application to typical structures, including the proto-
type buildings and their structural components.
Tsunami inundation can take the form of both a
rapidly rising tide or surge, or a broken bore. Bores
only develop under certain conditions but can result Tsunami loading differs
in higher lateral loads on structures and their ele- significantly from hurricane
storm surge loading!
ments. They are not known to form during drawdown,
so loading during outgoing flow need not consider Higher flow velocities make
bore conditions. tsunami loading considerably
more severe than storm surges
Although hydrodynamic loads occur during both
with equivalent flow depths.
tsunamis and hurricane storm surge, the higher flow
velocities during tsunamis make this loading condition
much more critical than during storm surge inunda-
tion. During hurricane storm conditions, wind-driven waves riding on top of the storm surge
can apply large impulsive lateral loads caused by wave breaking. These loads are considered
in ASCE 7-16, Chapter 5 and FEMA P-55 (2011) for coastal structures in hurricane zones. How-
ever, wave breaking is not considered in tsunami design because tsunami waves typically
break offshore where no land-based structures will be affected. For the design of offshore
structures in relatively shallow water depths or coastal defenses such as breakwaters and
other port structures, it may be necessary to consider tsunami wave breaking. This topic is
not addressed in ASCE 7-16, Chapter 6.
Theoretical studies show that solitary waves will typically break when the wave height
reaches approximately 78% of the still water depth (Weggel 1972, McCowan 1891). Given the
offshore wave heights in the ASCE 7-16 database, and illustrated in Figure G7-1 for Monterey,
California, it would be possible to apply standard shoaling theory to infer the wave height

111
@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


112 Tsunami Loads and Effects

increase as the tsunami moves into shallower water.


Tsunami waves typically One could then estimate the location at which the
break offshore, so coastal and
maximum anticipated tsunami wave would break,
land-based structures are not
required to consider wave- potentially loading structures located at that water
breaking loads. depth. Consideration must also be given to the pos-
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

sibility that a leading negative tsunami wave would


have reduced the coastal water depths owing to
drawdown. It is also likely that dispersion of the tsunami wave will result in a number of
smaller wave-breaking events rather than a single large wave break. These smaller waves
would break closer to the shoreline, in shallower water.
All these considerations are outside of the current requirements in the ASCE 7-16 stan-
dard, which is intended for application to coastal and on-land structures only.

10.1  Simplified Equivalent Uniform Lateral


Static Force
In many instances, the seismic and wind loading on a building will control over the tsunami
loading. This will be the case especially for taller structures in higher seismic hazard areas
with low tsunami flow depths. Buildings near the runup limit on the tsunami design zone
maps will have significantly lower tsunami loading than those close to the coastline.
To save the designer the task of performing a detailed calculation of hydrodynamic lat-
eral forces, a simplified equivalent uniform lateral static pressure method was developed by
the TLESC. As explained in ASCE 7-16 Commentary, Section C6.10.1, this method is based on
conservative assumptions of all parameters that influence lateral loading. Example G10-1
shows the application of this simplified method to one
of the prototype buildings introduced in Chapter 2. The
Applying the simplified equivalent static pressure is applied to the entire width
equivalent static lateral of the building perpendicular to both incoming and out-
approach may save the
going flow directions. The overall building lateral force–
designer time by avoiding
the more detailed loading resisting system and all individual structural members
evaluation. below a height of 1.3 hmax must be evaluated for this pres-
sure acting on the appropriate tributary width.
For incoming flow, the conservative assumption is
made that bore conditions exist, leading to increased lateral loading. The same expression
is used for outgoing flow, even though bores are not expected to form during drawdown.
This is because flow acceleration caused by topography and erosion during drawdown may
exceed the amplification factor assumed in development of the simplified approach.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Hydrodynamic Loads 113

Example G10-1. Simplified Equivalent Static Lateral Load.

In lieu of performing detailed hydrostatic and hydrodynamic analysis, ASCE 7, Equa-


tion  (6.10-1) provides a simplified but conservative estimate of the maximum lateral
load on the building. This approach is applied to the prototype buildings introduced
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

in Chapter 2. The building exterior is assumed to experience a uniform pressure, puw,


given by

puw = 1.25ITSUgshmax

where
ITSU = 1.0 (TRC II building),
gs = ksgsw = 1.1 × 64 = 70.4 pcf, and
hmax = 21.6 ft (for Monterey, California, location).

Therefore, fuw = 1.25 × 1.0 × 70.4 × 21.6 = 1901 psf


The total load on the building width, B = 254 ft, over a flow depth of 1.3 hmax
= 1.3 × 21.6 = 28.1 ft, is therefore

FTSU = 1,901 × 254 × 28.1/1,000 = 15,570 kips

For buildings designed for Seismic Design Category D, E, or F, this lateral load is
compared with 0.75 W0 Eh , where Eh is the seismic base shear from the building seismic
design. For the prototype buildings introduced in Chapter 2, the maximum seismic
base shear is Eh = 3,298 kips for the reinforced concrete shear wall building located in
Monterey, California, with a corresponding W0 = 2.5 (Table G2-3), so

FTSU= 0.75 × 2.5 × 3,298 = 6,184 kips < 15,570 kips.

Therefore, the detailed analysis for LC2 and LC3 must be performed as shown in
the following sections.

10.2  Detailed Hydrodynamic Lateral Forces


If the building or some of its components are not able to resist the simplified equivalent
uniform static lateral pressure in ASCE 7-16, Section 6.10.1, then all loading conditions in this
section must be considered to act both on the entire building and on individual structural
elements, as appropriate for the building being considered.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


114 Tsunami Loads and Effects

10.2.1  Overall Drag Force on Buildings and


Other Structures
As water flows around an obstruction, it applies unequal pressures on the front and back of
the obstruction and friction forces along the sides (Review G10-1). The resulting force on the
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

obstruction is referred to as the drag force given by ASCE 7-16 as

1
Fdx   s ITSU Cd Ccx B(hu 2 ) (6.10-2)
2
where
rs = ksrsw = 1.1 × 2.0 = 2.2 sl/ft3 [= 1.1 × 1,025 = 1,128 kg/m3],
ks = 1.1 is the fluid density factor to account for suspended soil and other smaller
flow-embedded objects,
rsw = 2.0 sl/ft3 [1,025 kg/m3] is the mass density of seawater,
ITSU = Tsunami importance factor from ASCE 7-16, Table 6.8-1,
Cd = Drag coefficient determined from ASCE 7-16, Table 6.10-1 for entire buildings and
ASCE 7-16, Table 6.10-2 for individual structural components,
Ccx = Closure coefficient representing the blockage to flow caused by structural mem-
bers and debris damming,
B = Overall width of the building or structural element,
h = Flow depth, and
u = Flow velocity.

For large objects such as a building, the magnitude of the unbalanced pressure differen-
tial between the front and back of the building will depend on the ratio between the building
width perpendicular to the flow, B, and the flow depth, hsx. For wide buildings it will take lon-
ger for the water level to equalize on opposite sides of the building, leading to large overall
drag on the building. This effect is captured by the drag coefficient, Cd, which changes based
on the B/hsx ratio as given in ASCE 7-16, Table 6.10-1 (repeated here as Table G10-1), which is
based on FEMA P-55, Coastal Construction Manual. Intermediate values are to be determined
by linear interpolation.
Hydrodynamic drag must be applied to all structural elements exposed to the tsunami
flow, including columns, beams, and walls. ASCE 7-16, Equation (6.10-3) is used to estimate
the closure coefficient, Ccx, which represents the percentage of the building elevation that
consists of structural elements expected to survive the tsunami flow. All nonstructural walls,
windows, and so forth are assumed to have been washed away by the time Load Case 2 is
reached. However, remaining nonstructural elements and accumulation of debris from adja-
cent buildings is included in the limiting value of Ccx given in ASCE 7-16, Section 6.8.7 and
illustrated in Section 8.7 of this guide.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Hydrodynamic Loads 115

Review G10-1. Hydrodynamic Drag.

Hydrodynamic drag is the resultant force experienced by an object when placed in


a fluid flow. The stagnation pressure at the leading edge of the object is p = 1/2ru2,
where ρ is the mass density of the fluid, and u is the velocity of the fluid. Numerous
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

laboratory experiments have shown that the overall drag force on the object, Fd, can be
determined by applying this stagnation pressure to the submerged projected area of
the object (width, b, by height, h), modified by an empirically determined drag coeffi-
cient. Therefore, the general equation for drag is expressed as

1 2 1
Fd   u bhCd   Cd b(hu 2 )
2 2

Drag coefficients, Cd, are given in ASCE 7, Table 6.10-2. They are low for objects
with low drag, such as a circular column (Cd = 1.2), and higher for more angular shapes,
such as square or rectangular columns, with one face oriented perpendicular to the
flow (Cd = 2.0). The value of Cd = 2.5 for a diamond-shaped column, pointed into the
flow, is based on the width of one face of the column and not the projected width of the
entire column. This is equivalent to Cd = 1.77 applied to the projected width for a square
column oriented at 45 degrees to the flow direction.
The term (hu2) in the drag equation is often referred to as the specific momentum
flux, or simply the momentum flux.

Table G10-1. Drag Coefficients for Rectilinear Structures.

Width to inundation depth ratio (B/hsx) Drag coefficient (Cd)


<12 1.25

16 1.3

26 1.4

36 1.5

60 1.75

100 1.8

>120 2.0

Source: ASCE 7-16, Table 6.10-1.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


116 Tsunami Loads and Effects

ASCE 7-16, Equation (6.10-3) for Ccx includes all structural columns and walls through-
out the building and not only at the front elevation. This is because water will flow around
leading columns and still apply drag loads to all interior and exterior columns on the back of
the structure. Engineering judgment can be used in cases such as elevator and stair shafts
where the front and back walls need not be summed because the hydrodynamic load will
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

be applied to the entire composite wall as a single element. Note that in Example G8-2 and
Example G8-3 only the front wall of each of the elevator or staircases is included in the Ccx
calculation because hydrodynamic load cannot develop on the leeward wall.
Experiments on fluid loading on bridge decks with multiple supporting girders has
shown that the full hydrodynamic load should be applied to the leading beam, but a reduced
lateral load is applied to subsequent beams (Douglass et al. 2006, Seiffert et al. 2015). In a
similar manner, fluid loads on the multiple beams supporting a floor slab will not all be equal.
The equivalent projected area of the beams is therefore represented by taking 1.5 times the
projected area of the slab edge and deepest beam perpendicular to the flow direction, as
illustrated in Example G10-2.
Example G10-3 shows the calculation of hydrodynamic drag forces on an entire build-
ing frame. The lateral loads are computed at each floor level so that they can be input into
an analysis program to evaluate the lateral force–resisting system for hydrodynamic loads.
However, for structural systems design as SDC D, E, or F, if the tsunami base shear, VTSU, which
is the sum of all elevated floor lateral loads, is less than 0.75 W0 Eh, as explained in Section
8.3.4, then no changes need to be made to the lateral force–resisting system. Therefore, it
would suffice to determine the overall lateral load on the building and compare this with the
seismic limit. Only if this test fails does one need to perform a lateral analysis to evaluate the
capacity of the lateral force–resisting system. This overall lateral force at ground level is also
required for evaluation of the foundation system, assuming that scour has already occurred,
as discussed in Chapter 12 of this guide. The lateral force–resisting system and foundation
system in the direction parallel to the flow must be evaluated for hydrodynamic loads using
the load cases and load combinations in ASCE 7-16, Sections 6.8.3.1 and 6.8.3.3. This load can
occur in both incoming and outgoing flow directions, so the lateral force–resisting system
must resist both directions equally.
For buildings oriented diagonal to the flow direction, both orthogonal lateral force–
resisting systems may participate proportionally to provide the tsunami load resistance as
shown in Example G10-4.

10.2.2  Drag Force on Component Elements


In addition to the hydrodynamic drag on the overall building frame given previously, each of the
structural elements submerged by the tsunami flow must be evaluated for the hydrodynamic
drag on the individual element. This drag force is given by ASCE 7-16, Equation (6.10-4) as

1
Fd   s ITSU Cd b(heu 2 ) (6.10-4)
2
@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Hydrodynamic Loads 117

Example G10-2. Beam Depth Used in Ccx.

The following figures show the appropriate value of beam depth, hbeam, to use when
calculating Abeam = B × hbeam for use in determining the closure coefficient, Ccx.
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

(a) Composite floor framing in a structural steel building.

(b) Pan joist concrete floor system.

(c) Beam-supported slab floor system.

(d) Column-supported two-way flat plate floor system.

(e) Flat slab with spandrel beams.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


118 Tsunami Loads and Effects

Example G10-3. Hydrodynamic Drag on Overall Building.

Consider the prototypical 7-story reinforced concrete residential building shown in


Figure G2-4. Assume a flow depth of 25 ft and a flow velocity of 15 ft/s, with the flow
direction perpendicular to the broad side of the building. Refer to Figure G2-4 for
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

dimensions of the structure.


Example G8-2 showed that the closure coefficient, Ccx, for the second floor
of this building was below the minimum of 0.7, so a value of Ccx=0.7 is used for the
hydrodynamic lateral load calculations. One could compute Ccx values for the other
floors or for the overall building frame below the inundation level, but they will not
vary significantly unless there is a change in the number of columns or walls, or a large
change in floor height.
The drag coefficient, Cd, is obtained from ASCE 7-16, Table 6.10-1 for a B/hsx ratio
of 254 ft/25 ft = 10.16 < 12; therefore, Cd = 1.25.
Because the flow level exceeds the third-floor level, the hydrodynamic drag on
the second and third floors will be computed individually, so that they could be input
as lateral loads in an analysis of the structural system. Then the total lateral load on
the building will be computed for design of the foundation system. Lateral load on the
lower half of the first floor will transfer directly to the foundation system, so it is not
included in the lateral load applied to the second floor.
1
The second-floor tsunami lateral load is given by Fdx   s ITSU Cd Ccx B (hu 2 )
2

where h = ½ (12 + 9 ft) = 10.5 ft.

1
So Fd 2 = × 2.2 ×1.0 ×1.25× 0.7 × 254 ×(10.5×152 ) / 1, 000 = 577.6 kips .
2

For the third-floor tsunami lateral load, h = 4.5 ft + (25 – 21 ft) = 8.5 ft, so Fd3 = 467.6 kips.
The tsunami base shear is therefore 577.6 + 467.6 = 1,045 kips.
The equivalent lateral load per unit height on the face of the building can be
determined by using h = 1 in the preceding equation to get Fdx = 55 kips/ft. This uni-
form distributed lateral load is applied over the lower 25 ft of the building as shown in
image (a) of the figure in Example G10-3.
The lateral load per floor can be determined using tributary areas as shown,
resulting in the floor lateral loads shown in image (b). Alternatively, the floor lateral
loads can also be determined by assuming that the exterior framing spans simply sup-
ported between the floor slabs, resulting in the floor lateral loads shown in image (c).
Finally, a continuous beam model could be used to more precisely determine the lat-
eral load at each floor level as shown in image (d).

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Hydrodynamic Loads 119

Example G10-3. Hydrodynamic Drag on Overall Building (Continued).


Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

As expected, these different approaches all result in the same foundation base
shear to be used for geotechnical design of the foundation system. However, there are
slight differences in the structural base shear, VTSU, determined at the midheight of
the ground floor columns. These differences are less than 1.5%, so any of the methods
would be acceptable to determine the structural base shear.
Because the prototype building was designed for SDC D, the structural base shear,
VTSU, is then compared with 0.75 W0 Eh, where Eh is the seismic base shear from the build-
ing seismic design. For this prototype building the seismic base shear is Eh = 3,298 kips
for the Monterey, California, location, with a corresponding W0 = 2.5 (Table G2-3), so

FTSU= 0.75 × 2.5 × 3,298 = 6,184 kips > 1,054 kips.

Therefore, the lateral force–resisting system is adequate without any strength-


ening for tsunami loads. The individual elements of the lateral force–resisting system
must still be evaluated as structural components following ASCE 7-16, Section 6.8.3.5,
which requires that structural components of the lateral force–resisting system be
designed for the forces resulting from the overall tsunami loads on the structural
system combined with any resultant actions caused by the tsunami pressures acting
locally on the individual structural component. This process is described more fully in
the worked example in the appendix of this guide.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


120 Tsunami Loads and Effects

Example G10-4. Hydrodynamic Drag on Overall Building for Diagonal Flow.

When the flow direction is not perpendicular to either face of the building, then the
orthogonal lateral framing systems will contribute proportionally to the resistance.
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

The hydrodynamic lateral load would be computed based on a building width per-
pendicular to the flow, B = 71 ft cos30 + 71 ft sin30 = 97 ft. The closure coefficient, Ccx,
should also be computed using the projected width of the columns, but is likely to be
similar to that determined for orthogonal flow.
With the flow direction as shown in the figure, the total hydrodynamic load, Fdx,
determined from ASCE 7-16, Equation (6.10-2) would be applied to the lateral force–
resisting systems in the two orthogonal directions as
(Fdx)x = Fdx cos30 in the x-direction, and (Fdx)y = Fdx sin30 in the y-direction.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Hydrodynamic Loads 121

In this expression the drag coefficient depends on the shape of the individual member
being considered. ASCE 7-16, Table 6.10-2 and Table G10-2 give appropriate Cd values for typ-
ical member shapes obtained from various established literature sources referenced in the
ASCE 7-16 Commentary. The effective width of the component, b, is the projected width per-
pendicular to the flow. Because of the potential for debris to accumulate against exterior
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

structural elements this projected width must be taken as the tributary width of the structural
element multiplied by the closure ratio given in ASCE 7-16, Section 6.8.7. Because the shape of
the debris is complex, it is required that a Cd value of 2.0 be used for all exterior components.
Example G10-5 shows the calculation of hydrodynamic drag force on a typical exterior column
in the prototype office building. Example G10-6 shows the calculation of hydrodynamic drag
force on an isolated exterior beam, whereas Example G10-7 shows the calculation of hydrody-
namic drag force on an interior column, which need not consider debris accumulation.

Table G10-2. Drag Coefficients for Structural Components.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


122 Tsunami Loads and Effects

Example G10-5. Exterior Column Drag (Including Debris Accumulation).

Consider an exterior column in the prototypical special moment-resisting frame


(SMRF) office building located in Monterey, California (see Chapter 2). Based on seis-
mic design, the column is 30 in. square. The columns are spaced 28 ft on center and
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

the debris closure coefficient is Ccx = 0.70. Therefore, the tributary width for hydrody-
namic drag on this exterior column is b = 0.70 × 28 = 19.6 ft.
Considering Load Case 2, with a flow velocity of u = 31.5 ft/s and a flow depth
exceeding the column height of 12.17 ft, the total hydrodynamic drag on this exterior
column is given by ASCE 7-16 as

1
Fd   s ITSU Cd b(heu 2 ) (6.10-4)
2

where
rs = ks × rsw = 1.1 × 2 = 2.2 sl/ft3,
ITSU = 1.0 (TRC II building),
Cd = 2.0 (required for member with debris accumulation),
b = Ccx × tributary width = 0.7 × 28 = 19.6 ft,
he = 12.17 ft (height of column from ground floor to soffit of second-floor beams),
and
u = 31.5 ft/s (maximum velocity for Load Case 2).

1
Therefore, Fd   2.2  1.0  2.0 

19.6 12.17  31.52
 521 kips

2 1, 000
Equivalent uniformly distributed load along the length of the columns is

w = 521/12.17 = 42.8 kips/ft

Example G10-6. Exterior Beam Drag (Including Debris Accumulation).

Consider an exterior beam in the SMRF office building located in Monterey, California.
Typically, the exterior beams are braced laterally by the floor slab, in which case there
is no need to check for lateral strength of the beams. However, for an isolated beam
that is not attached to the floor slab, the following derivation of lateral hydrodynamic
loads would apply.
The ground-floor height is 14 ft and the second-floor height is 12 ft. The maximum
flow depth is hmax = 21.5 ft.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Hydrodynamic Loads 123

For Load Case 3, the maximum flow depth, hmax = 21.5 ft, is combined with 1/3umax.
Because hmax exceeds the midheight of the second story (14 + 6 = 20 ft), the tributary
height of the exterior beam would be he = (14 + 12)/2 = 13 ft. With a debris closure
coefficient of Ccx = 0.70, the tributary height for hydrodynamic drag on this beam is he
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

= 0.70 × 13 = 9.1 ft. The flow velocity for Load Case 3 is umax/3 = 31.5/3 = 10.5 ft/s. The
hydrodynamic drag per unit length along this exterior beam is given by ASCE 7-16 as

1
Fd   s ITSU Cd b(heu 2 ) (6.10-4)
2
where
rs = ks × rsw = 1.1 × 2 = 2.2 sl/ft3,
ITSU = 1.0 (TRC II building),
Cd = 2.0 (required for member with debris accumulation),
b = 1 ft for load per unit length along the beam,
he = Ccx × tributary depth = 0.7 × (14 + 12)/2 = 9.1 ft, and
u = 31.5/3 = 10.5 ft/s (velocity for Load Case 3).

1
Therefore, Fd   2.2  1.0  2.0 

1 9.110.52 
 2.21 kips/ft
2 1, 000
This load must be applied as a uniformly distributed lateral load along the length
of the beam.
For Load Case 2, the flow depth is 2/3 hmax = 2/3 × 21.5 = 14.33 ft. The tributary
height for hydrodynamic drag on the beam is then he = 0.70 × (14.33 – 14/2) = 5.13 ft.
With a flow velocity of umax = 31.5 ft/s, the hydrodynamic drag per unit length along
this exterior beam is
1
Fd   s ITSU Cd b(heu 2 )
2
where
rs = ks × rsw = 1.1 × 2 = 2.2 sl/ft3,
ITSU = 1.0 (TRC II building),
Cd = 2.0 (required for column with debris accumulation),
b = 1 ft for load per unit length along the beam,
he = Ccx × tributary width = 0.7 × (14.33 – 7) = 5.13 ft, and
u = 31.5 ft/s (maximum velocity for Load Case 2).

F
Therefore, d 
1
 2. 2  1. 0  2. 0 

1 5.13  31.52 
 11.2 kips/ft
2 1, 000
This load is applied as a uniformly distributed lateral load along the length of the
beam and will control the design when compared with Load Case 3.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


124 Tsunami Loads and Effects

Example G10-7. Drag on Interior Column.

Interior structural members need not consider the effects of debris accumulation. For
example, for a typical 24 in. square interior column in the Monterey, California, SMRF
office building, the hydrodynamic drag force per unit height of the column is
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

1
Fd = rs ITSU Cd b(he u 2 )
2
where
rs = ks × rsw = 1.1 × 2 = 2.2 sl/ft3,
ITSU = 1.0 (TRC II building),
Cd = 2.0 [for square or rectangular column or beam from Table G10-2 (ASCE 7-16,
Table 6.10-2)],
b = 2 ft,
he = 1 ft, and
u = 31.5 ft/s (maximum velocity for Load Case 2).
1 2(1× 31.52 )
Therefore, Fd = × 2.2 ×1.0 × 2.0 × = 4.37 kips/ft
2 1, 000

This load is applied as a uniformly distributed load along the length of the column.

10.2.3  Tsunami Bore Loads on Vertical


Structural Components
As described in Section 6.4, tsunami bores often form as a tsunami approaches land. The bore
condition will survive until the Froude number drops below 1.0. Therefore, wherever bores
are to be considered according to ASCE 7-16, Section 6.6.4, and where the Froude number
exceeds 1.0, or where site-specific inundation modeling indicates bore conditions, increased
structural loading can result. The flowchart in Figure G10-1 shows this decision process.
Laboratory experiments performed in the large wave flume at the O. H. Hinsdale Wave
Research Laboratory at Oregon State University were used to derive an expression for the
maximum force applied to a solid wall by a broken bore (Robertson et al. 2013b). These exper-
iments involved generating a bore through wave breaking of a solitary wave. The resulting
hydraulic jump travels over standing water until it strikes a solid instrumented wall. The max-
imum pressure on the wall occurs during the initial bore strike on the wall (Figure G10-2);
however, the maximum total force on the wall occurs when the “splash-up” collapses and a
bore is formed moving away from the wall (Figure G10-3). At this point the pressure on the
wall is close to a triangular distribution.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Hydrodynamic Loads 125
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

Figure G10-1. Flowchart for bore loading on walls or columns.

Using the control volume identified in Figure G10-4, the following expression was devel-
oped for the maximum total force on the wall (Robertson et al. 2013b):

1 
Fw   sw  ghb2  h j v 2j  g 3  h j v j  3  (10-1)
1 4

2 
where
Fw = Maximum force on the wall,
rsw = Fluid density,
g = Acceleration due to gravity,
hb = hj + hs is the height of the bore, which includes the height of the incoming hydrau-
lic jump, hj, and the depth of the still water, hs, over which the jump is moving,
vj = Velocity of the hydraulic jump, which can be estimated using hydraulic jump
1 h 
theory as v j  ghb  b  1 , and
2  hs 
vs = 0 = Velocity of the standing water over which the hydraulic jump moves.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


126 Tsunami Loads and Effects
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

Figure G10-2. (Right) Maximum wall pressure and (lower left) corresponding total force as bore strikes
vertical wall. The red dot in the lower left image shows the total force on the wall at this instant in the
wall loading time history.

Figure G10-3. Video image, total force, and pressure distribution up the height of the wall at the point
of maximum lateral load applied to the wall by a bore.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Hydrodynamic Loads 127
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

Figure G10-4. Control volume immediately after bore strike on a solid wall used to develop expression
for total force on the wall.

This force acts as a triangular pressure distribution over a height of hp with a base pres-
sure of 2Fw/hp , where hp is the instantaneous ponding height at maximum load given by

 hj 
hp   0.25  1  hb  hr  (10-2)
 hs 

where

2
 v j hj 3
hr    (10-3)
 g
 

Comparing this theoretical expression with the experimentally measured force on the
wall during numerous tests at Oregon State University (OSU) results in the solid diamond
shapes in Figure G10-5. Perfect agreement is represented by the diagonal line, indicating
that this expression provides reasonable estimates of the total force on the wall.
Subsequent field surveys after the Tohoku tsunami in Japan identified a major structural
concrete wall at the Minami-Gamou wastewater treatment plant in Sendai that had been
damaged by a direct bore strike (Figure G10-6) (Chock et al. 2013b).
Video evidence was used to determine the bore flow depth and velocity at the time
of impact with the wall. The bore loading formula derived from the small-scale laboratory
experiments was applied to this full-scale field condition to determine the anticipated hydro-
dynamic loading on the wall. This load was then applied to a nonlinear numerical model of
the reinforced concrete wall, resulting in a yield line failure pattern and lateral load demand
matching what was observed in the field (Chock et al. 2013b, Robertson et al. 2013a). This
comparison confirmed that the loading expression derived from small-scale laboratory tests
can be applied to full-scale conditions.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


128 Tsunami Loads and Effects
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

Figure G10-5. Predicted versus experimental bore loading on wall.


Source: ASCE 7-16, Figure C6.10-1; Robertson et al. (2013b).

Figure G10-6. Reinforced concrete wall at the Minami-Gamou wastewater treatment plant damaged by
bore strike.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Hydrodynamic Loads 129

Also shown in Figure G10-5 are the results obtained using ASCE 7-16, Equation (6.10-
5b) (open diamond symbols). This equation gives a force that is 1.5 times the standard drag
force given by ASCE 7-16, Equation (6.10-5a). Figure G10-5 shows that this produces almost
as good a match with the experimental results, while also tending to be slightly conservative.
Therefore, the simpler Equation (6.10-5b) is used in ASCE 7-16 to determine the force applied
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

by a bore to a wall element, as demonstrated in Example G10-9.

Example G10-9. Tsunami Bore Loading on Structural Wall.

Consider the prototypical building with reinforced concrete shear walls shown in Fig-
ure G2-4. The wall width, b = 28 ft, is subject to bore strikes, and the Load Case 2 flow
depth hLC2 = 14.4 ft. Because hLC2 > b/3 = 9.33 ft, standard drag given by Equation (6.10-
5a) would apply for this flow depth, as illustrated in Part A. However, a bore of lower
height may occur as part of the tsunami inundation. The worst case would be if the
bore height were exactly equal to b/3 = 9.33 ft. This would require consideration of
bore impulsive loading given by Equation (6.10-5b), as shown in Part B.

Part A: Consider Load Case 2 as a non-bore flow of depth hLC2 = 14.4 ft and umax =
31.5 ft/s applied to the wall width b = 28 ft

Fw 
1
2

 s ITSU Cd b heu 2 
where
rs = 2.2 sl/ft3,
ITSU = 1.0 for Tsunami Risk Category II building,
Cd = 2.0 for rectangular element,
b = 28 wall width,
he = hLC2 = 14.4 ft, and
uLC2 = umax = 31.5 ft/s

So,
1
Fw   2.2 1.0  2.0 

28 14.4  31.52
 880 kips

2 1, 000

880
Therefore, the pressure on the wall is pw   2.18 kips/ft 2 acting from
14.4  28
the base of the wall to a height of 14.4 ft.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


130 Tsunami Loads and Effects

Example G10-9. Tsunami Bore Loading on Structural Wall (Continued).

Part B: Consider a bore of height hb = b/3 = 9.33 ft. Because b > 3hb, Equation (6.10-5b)
applies
3
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

Fw = rs ITSU Cd b (he u 2 )
4 bore

where he = hb = 9.33 ft, and ub = 28.4 ft/s (velocity corresponding to hb = 9.33 ft using
ASCE 7-16, Figure 6.8-1).

So,
3
Fw   2.2 1.0  2.0 

28 9.33  28.42 
 695 kips
4 1, 000

695
Therefore, the pressure on the wall is pw   2.66 kips/ft 2 acting from the
9.33  28
base of the wall to a height of 9.33 ft.
The wall must be checked in bending and shear for both of these conditions.

The results of these experiments confirmed earlier work by Ramsden (1993) on bore
loading on solid walls that showed a 1.5 increase in impulsive force compared with the
anticipated drag force. Árnason (2005) performed laboratory experiments on bores of dif-
ferent heights impacting a square column. For bore heights less than one-third the column
width, the 1.5 increase for the impulsive load is evident. However, for bore heights exceed-
ing one-third the column width, the impulsive load was not greater than the subsequent
drag force.
ASCE  7-16 therefore requires that Equation (6.10-5b) need only be considered for
cases in which the width of the column or wall element perpendicular to the flow is three
times the bore height (hb). If the width of the structural element is less than 3 hb, then the
standard drag force given by Equation (6.10-5a) applies. Strictly speaking, Section 6.10.2.3
only requires that this increased wall force be applied to “vertical structural components
that are wider than three times the inundation depth corresponding to Load Case 2 during
inflow.” Therefore, this increased force would only be applied if hLC2 = 2/3hmax < b/3. If hLC2>
b/3, then the standard drag force would apply [Equation (6.10-5a)]. However, because the
tsunami wave could consist of a series of bores of varying height, there is the possibility
that if the maximum flow depth, hmax, is greater than b/3, then a bore height matching b/3
could occur, so this condition should be considered as an additional load case. This con-
dition is demonstrated in Example G10-9. Note that Equation (6.10-5a, b) does not include
the debris accumulation coefficient, Ccx. This is because a bore strike is most likely to occur
at the leading edge of the tsunami, at which time debris will not yet have had time to accu-
mulate against the structure.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Hydrodynamic Loads 131

10.2.4  Hydrodynamic Load on Perforated Walls


The expressions developed in the previous section apply to solid walls impacted by a tsunami
bore. The expression included in ASCE 7-16 as Equation (6.10-5b) essentially represents an
effective drag coefficient, Cd, of 3.0 (1.5 times a Cd of 2.0). The drag coefficient for an indi-
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

vidual rectangular column is 2.0. Therefore, the tsunami bore loading on a perforated wall
(series of closely spaced columns or wall piers) will lie somewhere between these two values.
A series of experiments were performed in the O. H. Hinsdale Wave Research Laboratory
at OSU on perforated walls having different closure ratios as shown in Figure G10-7 (Robert-
son and Mohamed 2009). The center wall panel or column was instrumented to determine
the total hydrodynamic force on the column during the bore loading. Figure G10-8 shows
typical flow around two of these perforated wall configurations.
For closure ratios less than 20%, these experiments showed that there was no effect
on the maximum drag force on the center column. A drag coefficient of Cd = 2.0 was still
applicable. However, as the closure ratio increased from 20% to 100% there was an almost
linear increase in the effective drag force coefficient, primarily caused by increased buildup
of water on the upstream side of the specimen (visible in Figure G10-8, right). ASCE 7-16 pro-
vides for this variation in drag coefficient using Equation (6.10-6), which gives the force on a
perforated wall caused by a bore strike as

Fpw   0.4Ccx  0.6  Fw (10-4)

where Ccx is the closure coefficient given by the sum of the width of all columns or pier ele-
ments, over the total width of the wall.

Example G10-10. Hydrodynamic Load on Perforated Wall.

Consider the 28 ft wide shear wall used in Example G10-9. If the center 8 ft width of
this shear wall were constructed of lightweight breakaway wall (still with the fire rating
required for the elevator core), then the force on the remaining two 10 ft wide walls
would be given by
Fpw = (0.4Ccx + 0.6)Fw

where C=
20 3
= 0.714 , and Fw   2.2 1.0  2.0 

20 9.33  28.42 
 496 kips
cx
28 4 1, 000

Therefore, Fpw = (0.4 × 0.714 + 0.6) × 496 = 440 kips


440
The pressure on the perforated wall would then be p pw   2.36 kips/ft 2
9.33  20
compared with 2.66 kips/ft2 for the solid wall.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


132 Tsunami Loads and Effects
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

Figure G10-7. Solid and perforated wall arrangements for bore loading tests at OSU.

Figure G10-8. Multiple columns representing perforated walls with closure ratios of (left) 7% and
(right) 42%.

This expression is also illustrated in ASCE  7-16, Figure C6.10-2 in the commentary,
repeated here as Figure G10-9.

10.2.5  Walls Angled to the Flow


The wall forces given by ASCE 7-16, Equation (6.10-5a, b) are for the condition in which the
tsunami flow is perpendicular to the wall, that is, θ = 90 degrees. If the angle between the
flow direction and the wall is less than 90 degrees, then the force on the wall will be reduced
as given by ASCE 7-16 as

Fw  Fwsin 2 (6.10-7)

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Hydrodynamic Loads 133
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

Figure G10-9. Hydrodynamic drag transition from individual column to solid wall.
Source: ASCE 7-16, Figure C6.10-2.

where Fw is the force on the wall for flow perpendicular to the wall, from Equation (6.10-5a
or b); and θ is the angle between the flow direction and the wall.

10.3  Hydrodynamic Pressures Associated with Slabs


10.3.1  Flow Stagnation Pressure
When moving water enters and fills a sealed enclosure, the water inside the enclosure
becomes stationary. However, the exterior water is still applying the full stagnation pressure
at the entrance to the enclosure. As a result, stagnation pressure develops inside the enclo-
sure, given by ASCE 7-16 as
1
Pp   s ITSU u 2 (6.10-8)
2
where u is the water velocity outside the enclosure corresponding to the flow depth or load
case being considered.
Field observations during the Tohoku tsunami identified a number of failures caused
by stagnation pressure (Chock et al. 2013b). Three of these failures are described in Exam-
ple G10-12. Example G10-13 shows application of ASCE 7-16, Equation (6.10-8) to a mechan-
ical room surrounded by reinforced concrete shear walls in the prototypical residential
building (see Chapter 2).

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


134 Tsunami Loads and Effects

Example G10-11: Loading on Wall Angled to the Flow.

Part A: Consider the 28 ft wide shear wall used in Example G10-9. If the flow is orient-
ed such that the included angle between the flow direction and the wall axis is θ = 60
degrees, then the force on the wall (from Example G10-9, Part B) would reduce to
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

Fwθ = 695 × sin260 = 521 kips


521
The pressure on the wall would then be pw   2.00 kips/ft 2 compared
9.33  28
with 2.66 kips/ft2 for the perpendicular bore strike.

Part B: If the same 28 ft wide shear wall were oriented perpendicular to the shore-
line, that is, parallel to the flow direction, then theoretically there would be no lateral
hydrodynamic load on the wall. However, the ASCE 7-16 provisions require consider-
ation of the flow direction varying by 22.5 degrees either side of the shoreline normal
direction. For this condition, θ = 22.5 degrees, then the force on the wall (from Exam-
ple G10-9, Part B) would be

Fwθ = 695 × sin222.5 = 102 kips


102
The pressure on the wall would then be pw   0.39 kips/ft 2 , which is
9.33  28
much less than the 2.66 kips/ft2 for the perpendicular bore strike. However, structural
walls oriented parallel to the flow should still be checked for this condition.

10.3.2  Hydrodynamic Surge Uplift at


Horizontal Slabs
10.3.2.1  Slabs Submerged during Tsunami Inflow
Horizontal slabs submerged in tsunami flow will experience both upward and downward
hydrodynamic loads. Because slabs are generally designed for downward loads attributable
to gravity, they can usually accommodate the downward component of these hydrodynamic
forces. However, ASCE 7-16 requires that slabs submerged by tsunami flow be checked for an
uplift pressure of 20 psf in addition to buoyancy effects. Concrete floor slabs will generally
be heavy enough, even when submerged, to resist this uplift pressure (see Example G10-14).
However, lightweight concrete topping on metal deck floor systems in structural steel build-
ings may be more susceptible to uplift (see Example G10-15).

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Hydrodynamic Loads 135

Example G10-12. Flow Stagnation Example from the Tohoku Tsunami.

A cold-storage facility in Onagawa, Japan, was constructed of a reinforced concrete beam-


column frame with integral reinforced concrete wall and roof panels [Figure G10-10(a)].
The building was located approximately 200 m from the shoreline [Figure G10-10(b)].
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

A single double-door opening on the inland side of this building allowed water to enter
during drawdown and develop a stagnation condition. The internal pressure was suffi-
cient to fail the large wall panels on the near side of part (a), but not high enough to fail
the smaller wall panels to the left and back in part (a). Based on the stagnation pressure
equation it was estimated that the flow velocity during drawdown must have exceeded
5.5 m/s in order to fail the larger wall panels, but not greater than 7.5 m/s, which would
have caused blowout of the smaller panels. However, failure of the larger panels likely
reduced the subsequent internal pressurization, thereby preventing failure of the small-
er panels even if higher velocities were experienced.

(a)   (b)
Figure G10-10. (a) Reinforced concrete cold-storage building showing blowout wall panels and door
opening in the background; (b) Google Earth image showing building location relative to shoreline.
Source: (a) Courtesy of G. Chock; (b) Google Earth image, April 15, 2011.

A reinforced concrete cold-storage building located at the harbor in Onagawa


was effectively sealed except for a large double-door opening on the inland side of
the building [Figure G10-11(a)]. During drawdown, water entering this opening would
have filled the sealed space and resulted in the development of stagnation pressure
inside the building. This pressure was sufficient to fail a portion of the reinforced con-
crete slab roof and a number of the reinforced concrete wall panels. Reinforcing steel
samples were recovered and tested to estimate the failure strength of the wall panels
[Figure  G10-11(b)]. Based on the flow stagnation equation, it would have required a
flow velocity of 8.0 m/s during drawdown to cause the observed failures. Analysis of
survivor videos for the Onagawa harbor area indicates that the drawdown velocities
reached as high as 7.5 to 8.2 m/s (Chock et al. 2013a, Koshimura and Hayashi 2012).

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


136 Tsunami Loads and Effects
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

Figure G10-11. (a) Reinforced concrete cold-storage building with door opening on inland wall
showing roof and wall panel blowout caused by stagnation pressure; (b) sampling reinforcing
steel for strength tests.

A small reinforced concrete utility room at the Yuriage fishing boat harbor on the
Sendai plain had a single door opening facing the ocean, but no other window or door
openings. All walls and roof were reinforced concrete, effectively sealing the room except
for the single door opening. The building was completely overtopped during the Tohoku
tsunami. The internal stagnation pressure was sufficient to rupture the reinforced con-
crete at one of the back corners of the room, thereby relieving the internal pressure.
Samples of reinforcing steel were recovered for testing to estimate the tensile capacity
of the wall. The resulting analysis determined that the flow velocity at this location must
have reached 7.5 m/s or greater for the stagnation pressure to cause this failure.

Figure G10-12. (a) Interior view of utility closet looking through door opening toward the ocean;
(b) exterior view of rear of closet showing wall rupture; (c) Gary Chock assisting the author
recover reinforcing steel samples.
Source: (c) Courtesy of Tom Sawyer/ENR.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Hydrodynamic Loads 137

Example G10-13. Flow Stagnation.

Consider a condition in which one of the structural shear walls in the residential build-
ing shown in Figure G2-4 was used as a mechanical room or storage closet. This room
would be enclosed by reinforced concrete walls on all four sides and a reinforced con-
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

crete floor slab above and below the room. The walls are required for the gravity and
lateral load stability of the structure, so the walls must be designed to resist all tsunami
loads, including flow stagnation. Flow entering through one or both the doorways will
result in pressurization of the enclosure. Considering Load Case 2 with flow velocity of
31.5 ft/s, the stagnation pressure given by ASCE 7-16 would be

1
Pp   s ITSU u 2 (6.10-8)
2

where
rs = 2.2 sl/ft3,
ITSU = 1.0 for a Tsunami Risk Category II building, and
u = 31.5 ft/s for Load Case 2.

1
Therefore, Pp   2.2 1.0  31.52  1, 091 psf
2

All four walls and roof slabs must be designed to resist the bending moments and
shear forces resulting from this outward pressure.

(a) Plan view of enclosed mechanical/storage room in residential building; (b) section through
mechanical/storage room.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


138 Tsunami Loads and Effects

Example G10-14. Slab Uplift Example 1.

Consider a 6 in. thick reinforced concrete flat slab submerged in tsunami flow.
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

Hydrodynamic uplift required by ASCE 7-16, Section 6.10.3.2.1 is 20 psf.


Slab buoyancy is 6/12 × 1.1 × 64 pcf = 35.2 psf.
Slab self-weight is 6/12 × 150 pcf = 75 psf.
The appropriate load combination is

0.9D + FTSU = 0.9 × 75 – (35.2 + 20) = 12.3 psf “downward” load.

The uplift forces do not exceed the deadweight, so the slab will not need any addi-
tional design for the uplift forces.

Example G10-15. Slab Uplift Example 2.

Consider a 3 in. thick lightweight concrete topping over a 3 in. metal deck as shown in
the following:

Hydrodynamic uplift required by ASCE 7-16, Section 6.10.3.2.1 is 20 psf.


Slab buoyancy is 6/12 × 1.1 × 64 pcf = 35.2 psf, including the effect of air trapped
below the metal deck.
Slab self-weight is 4.5/12 × 120 pcf plus metal deck weight of 2 psf = 47 psf.
The appropriate load combination is

0.9D + FTSU = 0.9 × 47 – (35.2 + 20) = -12.9 psf “upward” load.

The uplift forces exceed the deadweight, so the slab will need to be checked for
negative bending, and the connections to the supporting beams must be able to resist
this uplift force. The supporting beams, columns, and foundations will also have to
resist this uplift force.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Hydrodynamic Loads 139

10.3.2.2  Slabs over Sloping Grade


If an elevated structural floor is located over a sloping grade, with angle φ greater than
10 degrees, then the vertical component of the flow velocity will impart an upward pressure
on the soffit of the slab. This will occur during tsunami inflow if the grade is sloping upward
away from the coast (Figure G10-11), and will occur during drawdown if the grade is sloping
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

downward away from the coast.


This uplift pressure is given by ASCE 7-16, Equation (6.10-9) as

Pu  1.5 ITSU  s uv2  20 psf (6.10-9)

where
u = Horizontal flow velocity corresponding to a water depth equal to or greater than
the elevation of the soffit of the slab, hss,
uv = utanφ, and
φ = Average slope of grade plane beneath the slab.

The 20 psf minimum limit is based on ASCE 7-16, Section 6.10.3.2.1 and corresponds to a
flow velocity of 14 ft/s (4.3 m/s) over a slope of 10 degrees. Example G10-16 shows an exam-
ple calculation of this hydrodynamic uplift.
This expression is based purely on the direction of fluid moving up the slope and does
not account for the potential pressure increase as the opening between the ground and
elevated floor slab becomes restricted at the back of the building. Additional experimenta-
tion or computational modeling is required to determine the effect of this narrowing of the
flow channel, and the corresponding acceleration of the flow, which may tend to reduce the
pressure. If the slope is such that there is no gap at the back of the building, then the uplift
pressures will tend toward those given in ASCE 7-16, Section 6.10.3.3.
This hydrodynamic uplift load must be combined with any buoyancy because of water
displaced by the structural floor, including trapped air between the beams, joists, and gird-
ers, as well as potential uplift owing to delayed breakaway of the nonstructural exterior walls
around the upper floor, as illustrated in Figure G10-13. ASCE 7-16 provides no guidance on
the realistic estimation of breakaway strength of nonstructural exterior walls, which can
often be substantially greater than the required strength for wind. Example G9-2 in Sec-
tion 9.1 demonstrates how this buoyancy effect can be estimated.

10.3.3  Tsunami Bore Flow Entrapped in


Structural Wall-Slab Recesses
Laboratory experiments have shown that when a tsunami bore becomes entrapped below
a structural slab and its progress is prevented by a structural wall, then large pressures
develop on the slab and wall (Ge and Robertson 2010). Similar pressures also develop on the

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


140 Tsunami Loads and Effects

Example G10-16. Slab Over Sloping Grade Uplift.

Consider the 7-story residential building shown in Figure G2-4, but located over a
grade sloping at 20 degrees. For Load Case 2, the flow depth is 14.4 ft, which exceeds
the height of the slab soffit (12 ft – 8 in./12 = 11.33 ft). The corresponding flow velocity
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

is umax = 31.5 ft/s, so this represents the controlling condition for hydrodynamic uplift
pressure on the soffit of the second-floor slab given by ASCE 7-16 as

Pu = 1.5ITSUrsuv2 ≥ 20 psf (6.10-9)

where
ITSU = 1.0 for TRC II building,
rs = ksrsw = 1.1 × 2.0 = 2.2 sl/ft3, and
uv = utanφ = 31.5 × tan 20°= 11.5 ft/s.

Therefore, Pu = 1.5 × 1.0 × 2.2 × 11.52 = 434 psf ≥ 20 psf.

The potential additional uplift caused by buoyancy will depend on the failure
strength of the nonstructural exterior walls at the second level. Example G9-2 shows
how this buoyancy of a submerged elevated floor can be determined. These two uplift
effects must be combined when designing the floor system. Overall uplift of the build-
ing must also be checked for this combination of uplift loading.

floor below, but if this is a slab-on-grade, then these pressures will be of no concern. If the
slab below is itself an elevated slab, it will feel the large downward pressure, but may already
be inundated by water below the slab, which may also be pressurized by the incoming flow.
Experiments were performed at the O. H. Hinsdale Wave Research Laboratory at Ore-
gon State University in both the large wave flume (LWF) (Figure G10-14 and Figure G10-15)
and tsunami wave basin (TWB) (Riggs et al. 2005). These experiments utilized a solitary wave
breaking over a simulated fringing reef to generate a bore, which then impacted the slab-
wall test specimen. Tests were also performed using a dam-break bore generator in a wave
flume at the University of Hawai‘i at Manoa (Ge and Robertson 2010). The experimental setup
used at UH Manoa was identical to that used in the TWB at OSU. The experiments at UH
Manoa and in the TWB at OSU were at approximately 1:10 scale, whereas the experiments in
the LWF at OSU were at approximately 1:5 scale.
Numerous waves were run at each facility with varying bore heights, standing water
depths, and slab elevations. Figure G10-15 shows one of the waves producing the largest slab
uplift load in the LWF at OSU. The pressure on the wall at the back of the slab is approximately
equal to the uplift pressure on the soffit of the slab close to the wall. Moving away from the
wall, the slab uplift pressure drops to about half the maximum value. Figure G10-16 shows
the average uplift pressure on the soffit of the slab for all the test results plotted relative to

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Hydrodynamic Loads 141
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

Figure G10-13. Uplift on elevated floor owing to sloping grade and buoyancy.

Figure G10-14. Slab uplift experiment in large wave flume at O. H. Hinsdale Wave Research Laboratory, OSU.
Source: Riggs et al. (2005).

the ratio between slab height, hs, and bore height, h. The maximum uplift pressures were
recorded when the bore height exceeds two-thirds of the slab height, that is, h ≥ 2/3hs,
or hs/h ≤ 1.5. As the slab height increases relative to the bore height, the uplift pressure
decreases.
An envelope enclosing 90% of the test results is shown by the black dashed line in
Figure G10-16. The maximum plateau on this envelope is at 266 psf (12.77 kPa). Based on

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


142 Tsunami Loads and Effects
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

Figure G10-15. Video and data capture at peak uplift load on slab during experiment shown in
Figure G10-14: (a) video capture at peak load; (b) total slab uplift force time history; (c) uplift pressure
distribution on soffit of slab; (d) pressure distribution on wall behind slab.
Source: Riggs et al. (2005).

the observation that half of the slab was subjected to higher pressures (close to the wall),
and the other half was subjected to 50% lower pressures (Figure G10-15c), ASCE 7-16, Sec-
tion 6.10.3.3.1 requires that portion of the slab closest to the wall be designed for uplift pres-
sure of 350 psf (16.76 kPa), which is approximately 1.333 × 266. The adjacent section of the
slab must be designed for 350/2 = 175 psf (8.38 kPa), which is approximately 0.667 × 266. The
average pressure on the slab would then be (350 + 175)/2 = 263 psf (12.62 kPa), which is close
to the upper plateau of the average pressure envelope shown in Figure G10-16.
If the bore height is less than two-thirds of the slab height, then these pressures reduce
linearly as illustrated by the dashed and dotted red lines in Figure G10-16. A minimum pres-
sure of 30 psf (1.44 kPa) is required regardless of the slab height relative to bore height. Based
on ASCE 7-16, Equation (6.10-10), this minimum pressure applies whenever hs/h ≥ 3.5, that is,
the slab height is 3.5 times the bore height.
Tsunami waves often occur as a series of bores or broken wave fronts. It is difficult to
predict the number or size of bores that will form as part of the incoming tsunami; how-
ever, for the purposes of computing the pressure that develops in a wall-slab recess, it is

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Hydrodynamic Loads 143
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

Figure G10-16. Average uplift pressure on the slab specimens during slab-wall experiments at two
different scales in three different wave flumes. A 90% envelope is shown by a black dashed line, and
the ASCE 7-16 loading expressions for elevated slabs are shown in red dashed lines.

conservative to assume the largest possible bore, that is, hbore = hmax. Any bore with a lesser
height will develop the same or lower pressures.
Failures caused by this type of loading have been observed in pile-supported wharves
adjacent to solid bulkheads and wharves over sloping riprap, as well as elevated floors in
buildings. During the 2004 Indian Ocean tsunami the precast slabs forming the deck of a
wharf at Kao Lak Harbor, Thailand, were lifted off the supporting beams (Figure G10-17). It is
unknown if there was any structural connection between the deck panels and the supporting
beams, but the uplift pressure was clearly able to lift the self-weight of the concrete panels.
Similar damage was observed in a shopping center in Patong Beach, Thailand, during the
same event, when precast floor slabs were lifted off the supporting beams (Figure G10-18).
During the September 2009 tsunami that impacted Samoa, a new wharf structure was
destroyed by the incoming wave because of pressure that developed between the slop-
ing grade and wharf deck (Figure G10-19). A small PVC pipe opening in the deck may have
been effective at releasing air trapped ahead of the water, but it was ineffective at releas-
ing the water pressure. The resulting uplift of the wharf deck allowed the water to pass
through the back of the wharf and uplift a slab-on-grade covering the adjacent apron area
[Figure G10-20(b and c)].

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


144 Tsunami Loads and Effects
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

Figure G10-17. Harbor wharf uplift failures at Kao Lak Harbor, Thailand, during the 2004 Indian Ocean
tsunami.
Source: CAEE (2005).

During the Tohoku tsunami there were numerous examples of wharf uplift caused by
the incoming flow. Many wharves had “breakaway panels” between the pile-supported wharf
and the grade-supported apron slab. These breakaway panels were lifted by the pressure
that developed below the wharf. Based on the weight of the concrete panels and the sur-
face area exposed to uplift pressure, Chock et al. (2013b) show that the uplift pressure must
have been in the range of 167 to 250 psf (8 to 12 kPa) to cause these panels to lift out of posi-
tion. Although there is no information about whether these wharves were hit by tsunami
bores or surging flow, nor is there information on the flow depth that might have caused the
observed damage, this range of pressures appears reasonable compared with the experi-
mental results shown in Figure G10-16, where this pressure range is shown as a shaded zone.
If these wharves did not have the breakaway panels, it is likely that the pressures would have
been higher, thereby justifying the higher envelope based on the experimental data.

10.3.3.1  Pressure Load in Structural Wall-Slab Recesses


Pressure load in structural wall-slab recesses needs only be considered if bore conditions
exist at the site (see Section 6.4). All the experimental data available at the time ASCE 7-16,
Chapter 6 was developed related to bore loading conditions. The lack of experimental results
for tsunami surge loading on wall-slab recesses meant that the code requirements are limited
to bore conditions. However, assuming that there is no pressurization of wall-slab recesses
under surge conditions is clearly unconservative, so it would be prudent to consider the
bore loading conditions as an upper bound for surge loading until such time as the neces-
sary experimental and/or field data are available to develop a reliable loading expression for
surge conditions.
Figure G10-16 showed the relationship between pressure developed by a bore in a wall-
slab recess and the ratio between the height of the slab, hs, and the height of the bore, h.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Hydrodynamic Loads 145
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

Figure G10-18. Second-floor slab uplift failure at shopping center in Patong Beach, Thailand, during the
2004 Indian Ocean tsunami.
Source: CAEE (2005).

Figure G10-19. Damage to new inter-island ferry dock structure in Malaela, Samoa (S 14.02553; W
171.42435): (a) view of the wharf from the ocean; (b) section of the wharf with blowout at the back;
(c) view through the blowout at the back of the wharf; (d) original cavity between the wharf deck and
the concrete covered sloping grade; a small PVC hole is visible in the deck slab.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


146 Tsunami Loads and Effects
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

Figure G10-20. Tohoku tsunami examples of wharf access panel uplift caused by tsunami hydrodynamic
loads: (a) Kessenuma Port (N 38.896; W 141.584); (b) Ofunato Port (N 39.0649; W 141.7244); (c) Onagawa
Port (N 38.4412; W 141.4498); (d) Yuriage Port (N 38.1725; W 140.9564).

When the bore height exceeds 2/3hs (or hs/h ≤ 1.5) then Pu = 350 psf. When the bore height is
less than 2/3hs, then the slab uplift pressure reduces following ASCE 7-16, Equation (6.10-10).
For values of hs/h ≥ 3.5, Equation (6.10-10) gives values of uplift pressure less than 30 psf, so
the minimum value of 30 psf applies. It is conservative to assume that the bore height equals
the maximum flow depth, hmax.
Figure G10-21 provides a flowchart to determine which loading conditions apply for a
given bore loading condition on a wall-slab recess.
Based on the preceding experimental and field data, ASCE 7-16, Section 6.10.3.3.1 requires
that for a bore height greater than two-thirds of the clear height to the bottom of the slab, hs,
then the slab within a distance hs from the wall must be designed for 350 psf uplift. Beyond a
distance of hs, the uplift reduces to 175 psf. If the wall has a width of lw and the water is able to
flow around the ends of the wall, then the uplift pressure is further reduced to 30 psf beyond
a distance of hs + lw from the wall. This loading is shown schematically in Figure G10-22. If
water is not able to flow around the ends of the wall, then the 175 psf uplift applies to the
entire slab, except within hs of the wall where the uplift is 350 psf. Note that the structural

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Hydrodynamic Loads 147
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

Figure G10-21. Flowchart for bore loading on wall-slab recess.

wall blocking the bore flow must also be designed for 350 psf if it is to survive the bore load-
ing. In the case shown in Figure G10-22, the back wall is a retaining wall, so passive pressure
developed in the soil behind the wall will contribute to resisting this lateral pressure. Exam-
ple G10-17 shows how these pressures are determined for a building with structural walls
preventing flow through the ground floor level.

10.3.3.2  Reduction of Load with Inundation Depth


The high uplift pressures provided in ASCE 7-16, Section 6.10.3.3.1 only apply if the clear slab
height to bore height ratio is less than 1.5, that is, the bore height, h, is greater than two-
thirds of the clear slab height, hs. If the bore is lower than two-thirds times the clear story
height, then the uplift force reduces linearly as given by ASCE 7-16, Equation (6.10-10) and
shown in Figure G10-16. The minimum uplift pressure is 30 psf, which applies for all cases

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


148 Tsunami Loads and Effects

Example G10-17. Wall-Slab Recess Pressure.

A building is exposed to bore conditions with a structural wall blocking the flow at the
back and sides of the ground floor level as shown in the figure. Based on ASCE 7-16,
Section 6.6.4, bore conditions are anticipated at the site. The energy grade line analysis
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

gives the maximum flow depth at the site as hmax = 24 ft, and the maximum velocity as
umax = 35 ft/s.

The flowchart in Figure G10-21 provides the following guidance:

1. Do bore conditions exist? Yes


2. Is Fr>1.0? Yes

u 35
For Load Case 2, Fr = = = 1.54 > 1.0 .
gh 32.2 × 2 / 3× 24

3. Is the slab height greater than 3.5h? No

hs = 17.25 ft  3.5× h = 3.5× 2 / 3× 24 = 56 ft

4. Is h greater than two-thirds slab height? Yes

2 2
h = 2 / 3× 24 ft = 16 ft > × hs = ×17.25 = 11.5 ft
3 3

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Hydrodynamic Loads 149

Therefore, the wall and slab within hs = 17.25 ft of the wall must be designed for
350 psf uplift pressure, whereas the rest of the slab must be designed for 175 psf uplift
pressure. Because of the side walls, flow cannot pass around the ends of the back wall;
therefore, the pressure never drops to 30 psf. The side walls will feel the same pres-
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

sures as the adjacent slab as shown in the plan view in the figure.
Because the slab will be submerged, the self-weight will be reduced by buoyancy.
Therefore, the net uplift pressure for which the slab and supporting beams adjacent to
the wall must be designed is

Pnet  FTSU  0.9 D  350  0.9  ts   conc  k s sw 

9
 350  0.9  150  1.1 64   296.3 psf
12

9
The net uplift reduces to Pnet  175  0.9  150  1.1 64   121.3 psf for the slab
12
further than 17.25 ft from the back wall.
These are large net upward pressures that will require flexural reinforcement
opposite to what is required for gravity loads. In addition, it is necessary to check for
upward punching shear at all slab-column connections, upward shear on beams, and
upward “beam” shear in the slab at the face of the walls and beams.

where hs/h ≥ 3.5, based on ASCE 7-16, Equation (6.10-10). Example G10-18 shows how these
pressures are determined for a building with a structural wall preventing flow through the
ground-floor level.

10.3.3.3  Reduction of Load for Wall Openings


Experiments performed in the TWB in the O. H. Hinsdale Laboratory at OSU showed that the
uplift pressure described in Section 10.3.3.2 reduces if there are openings in the wall at the
back of the slab-wall recess (Riggs et al. 2005). These experiments used a series of piers to
represent a “perforated wall” as shown in Figure G10-23. The closure coefficient, Ccx, of the
perforated wall is defined as the ratio of the solid area of the wall to the total inundated area
of the vertical plane of the wall. In the experiments this ratio was varied from 7% to 63% by
changing the width of the three piers shown in Figure G10-23. The resulting average uplift
pressure on the slab is compared with the solid wall condition (Ccx = 100%) in Figure G10-24.
The error bars indicate the range of experimental results for each closure coefficient, and
the black squares are the mean values. It was concluded that the uplift pressure on the slab
reduces proportional to the closure coefficient, resulting in Equation (6.10-11) in ASCE 7-16,

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


150 Tsunami Loads and Effects
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

Figure G10-22. Schematic showing uplift pressures on elevated slab when a bore is prevented from
passing through the building because of a structural wall. Conservatively the height of the bore, hb, is
taken equal to the maximum flow depth, hmax.

Figure G10-23. Slab-wall experiments with perforated wall (slab shown as transparent for clarity).

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Hydrodynamic Loads 151

Example G10-18. Wall-Slab Recess Pressure.

The same building used in Example G10-17 is exposed to bore conditions with a struc-
tural wall blocking the flow at the back of the ground-floor level as shown in the fig-
ure. Based on ASCE 7-16, Section 6.6.4, bore conditions are anticipated at the site. The
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

energy grade line analysis gives the maximum flow depth at the site as hmax = 15 ft, and
the maximum velocity as umax = 20 ft/s. Load Case 2 flow depth is h = 2/3 hmax = 10 ft.

The flowchart in Figure G10-21 provides the following guidance:

1. Do bore conditions exist? Yes


2. Is Fr > 1.0? Yes
u 20
The Froude number at the site is Fr    1.11  1.0 .
gh 32.2 10
3. Is the slab height greater than 3.5h? No

hs = 17.25 ft  3.5× h = 3.5×10 = 35 ft

4. Is h greater than two-thirds slab height? No

2 2
hmax  10 ft   hs  17.25  11.5 ft
3 3

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


152 Tsunami Loads and Effects

Therefore, the wall and slab within hs = 17.25 ft of the wall must be designed for the
uplift pressure given by ASCE 7-16, Equation (6.10-10), and the rest of the slab must be
designed for half of this pressure

 h   17.25 
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

Pu  ITSU  590  160 s   1.0  590  160 10   314 psf


 h   

This uplift pressure reduces to Pu = 0.5 × 314 = 157 psf for the slab further than
17.25 ft from the back wall. Because there are no side walls preventing the flow from
passing around the ends of the back wall, this pressure of 157 psf can be reduced to
30 psf at a distance hs + lw = 17.25 + 50 = 67.25 ft away from the wall. This does not occur
in the given building because it exceeds the building length.

Pur = CcxPu, where Pu is given by Equation (6.10-10) for the solid wall condition. Example G10-19
shows the application of this provision to determine the hydrodynamic pressures in a build-
ing with a perforated structural wall restricting flow through the ground-floor level.

10.3.3.4  Reduction of Load for Slab Openings


During the Tohoku tsunami, concrete access panels and steel grating spanning between the
pile-supported wharves and the soil-supported dock allowed for relief of some of the pressure
developed below the wharf (Figure G10-20). Experiments in a wave flume at the University

Figure G10-24. Change in slab uplift pressure with varying closure coefficient for the back wall in a
slab-wall recess subjected to bore loading.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Hydrodynamic Loads 153

Example G10-19. Wall-Slab Recess with Perforated Wall.

The same building used in Example G10-17 is exposed to the same bore conditions, but
with a series of three structural piers, each 10 ft wide, blocking the flow at the back of
the ground floor level as shown in the figure. The closure coefficient is therefore given
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

by Ccx = (3 × 10)/50 = 0.6. Therefore, the pressures will be 60% of those found for the
solid wall in Example G10-17. For the wall piers and the slab within hs = 17.25 ft of the
wall, Pur = 0.6 × 314 = 188 psf. For the rest of the slab, Pur = 0.6 × 157 = 94 psf as shown
in the figure.

of Hawai‘i at Manoa using a bore generated by dam break confirmed these results (Takakura
and Robertson 2010). The test setup for these experiments is shown in Figure G10-25. The
clear slab height, hs, and gap width, wg, were varied to study the effect on the average slab
uplift pressure. Figure G10-26 shows how the uplift on the slab reduces with the increased
gap width. Two slab heights were considered, and a number of different bore scenarios were
generated by the dam break. The curves are plotted through the mean test results, with the
range of results shown by the error bars. When the gap width equals half the slab height, the
uplift is reduced by more than 50%. The effect is less dramatic as the gap increases further.
ASCE 7-16, Equations (6.10-13) and (6.10-14) are based on these data and shown by the dashed
lines in Figure G10-26. Example G10-20 shows how these pressures are determined for a
building with a structural wall preventing flow through the ground-floor level, but a break-
away panel provided in the second-level floor slab.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


154 Tsunami Loads and Effects
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

Figure G10-25. Test setup to study the effect of a slab opening on the uplift pressure when a bore enters
a slab-wall recess.

Figure G10-26. Effect of slab opening width on average uplift pressure.

10.3.3.5  Reduction in Load for Tsunami Breakaway Wall


If the wall blocking flow through the building is designed as a tsunami breakaway wall, then the
maximum pressure that needs to be considered on the slab adjacent to the wall is equal to the
pressure required to reach the nominal shear force necessary to fail the breakaway wall. This fail-
ure is usually governed by separation of the wall from the slab at top or bottom of the wall. Esti-
mating this nominal shear force should take into account the likely overstrength of the connection
fasteners and hardware involved in making the connection between the breakaway wall and slab.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Hydrodynamic Loads 155

Example G10-20. Wall-Slab Recess with Breakaway Slab.

The same building used in Example G10-17 is exposed to the same bore conditions. The
slab is constructed with a loose breakaway panel adjacent to the back wall as shown in
the figure. The clear opening for the gap is 10 ft wide, and the gap runs the full width
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

of the building.

Because wg = 10 ft is greater than 0.5hs = 0.5 × 17.25 = 8.625 ft, ASCE 7-16, Equation
(6.10-14) is used to determine the effect of the breakaway slab as

wg 10
Cbs  0.56  0.12  0.56  0.12  0.49
hs 17.25

Therefore, the uplift pressure on the slab soffit given by ASCE 7-16, Equation (6.10-
12) is

Pur = CbsPu = 0.49 × 350 = 171.5 psf

This pressure is applied to the wall and the slab within hs = 17.25 ft of the wall
(including the breakaway panel). For the rest of the slab, Pur = 0.5 × 171.5 = 85.8 psf as
shown in the figure.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


156 Tsunami Loads and Effects

To verify that this pressure is sufficient to lift the breakaway panel, we must check
that it exceeds the dead weight of the panel

9
wD   concV  150  11  112.5 psf
12
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/18/20. Copyright ASCE. For personal use only; all rights reserved.

Because 171.5 psf is greater than 112.5 psf, the slab will lift and release the pressure
below the slab. If the slab weight was greater than 171.5 psf, then the wall-slab recess
pressure would be increased to match the slab weight. Alternatively, the slab weight
could be reduced by using lightweight concrete or a thinner slab for the breakaway
panel.
Care should be taken to design the columns at the second level to resist impact
from these breakaway slabs following the procedure in ASCE 7-16, Section 6.11.4, “Sub-
merged Tumbling Boulder and Concrete Debris,” even if these columns would not nor-
mally be exposed to debris impact.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


11
Debris Impact Loads
Downloaded from ascelibrary.org by La Trobe University on 10/20/20. Copyright ASCE. For personal use only; all rights reserved.

During coastal inundation, large amounts of floating and rolling debris are transported by
the tsunami. During the 2011 Tohoku tsunami, the debris consisted of everything from small
plastic items to very large ships. This section addresses the impact loads that need to be
considered in structural design owing to the potential for debris striking a structural mem-
ber. Debris impact strikes need not be considered if the flow depth is less than 3 ft since only
smaller items will float under this condition.
ASCE 7-16 considers five specific debris items, which are considered representative of
the more damaging types of debris found in a typical tsunami flow. These debris types are

1. Wood poles and logs, which are generated when the flow causes failure of trees, power
poles, and large timber structures.
2. Floating vehicles in the form of cars and trucks that are buoyant until water is able to
seep into the enclosed passenger compartment.
3. Submerged and tumbling boulder and concrete debris represents large dense items
that do not float but are rolled along the ground by the tsunami. Flow conditions have
to be adequate to initiate movement of these objects, but once they are moving, they
can induce significant impact damage just above grade level. These impacts need not be
considered if the flow depth at the site is less than 6 ft.
4. Shipping containers are universally used to transport goods through ports worldwide.
Large quantities of these containers both empty and with contents are stored at most
cargo ports, making them susceptible to flotation during a tsunami. Limits on the max-
imum weight of a loaded shipping container are set so that they can still be handled by
the port equipment. This total weight is about one-third of the buoyant force for the
closed container, so all containers will float if they are closed and subjected to more
than about 3 ft of water (Review G11-1).
5. Large ships and barges represent a major impact threat to buildings and other struc-
tures adjacent to ports. However, they require significant draft before they will float.
ASCE 7-16 only requires that these impacts be considered in the vicinity of ports and
harbors if the flow depth at the site exceeds 12 ft.

157
@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


158 Tsunami Loads and Effects

Review G11-1. Shipping Container Buoyancy.

A standard 20 ft shipping container has a self-weight of 5,400 lb. Maximum total weight
limit for a standard 20 ft container is 53,000 lb. The container dimensions are 20 ft
long by 8 ft wide by 8.5 ft high (or 9.5 ft for high cube).
For an empty 20 ft container to float, FBuoyancy ≥ WContainer, or, gsV = 70.4(20 × 8 × d)
≥ 5,400 ⸫ d ≥ 0.48 ft will cause the empty container to float.
Downloaded from ascelibrary.org by La Trobe University on 10/20/20. Copyright ASCE. For personal use only; all rights reserved.

For a sealed fully loaded container to float, gsV = 70.4(20 × 8 × d) ≥ 53,000 ⸫ d ≥


4.7 ft will cause the container to float.
“Heavy tested” 20 ft containers have a maximum weight of 67,200 lb and therefore
require 6.0 ft of water to float.
Although the self-weight (8,800 lb) and maximum weight (67,200 lb) of a 40 ft
container are larger than the standard 20 ft container, so is the displaced volume. The
result is that an empty 40 ft container will float in 0.39 ft of water, whereas a fully load-
ed 40 ft container will float in 3.0 ft of water.
ASCE 7-16 only requires consideration of shipping container debris impact when
the inundation depth exceeds 3 ft (Section 6.11).

Figure G11-1 provides a flowchart that identifies the types of debris that must be consid-
ered at a particular site. Debris impacts can occur during both incoming and outgoing flow,
but their effects need only be considered when designing perimeter gravity load–carrying
structural components. Damage to a gravity load–carrying structural component such as a
column can lead to progressive collapse of a portion or all of the building, potentially result-
ing in loss of life for those who might have sought refuge in the building. Interior structural
members are considered to be protected from debris by the perimeter members, or at least
the debris velocity is assumed to have decreased significantly if the debris is able to enter the
interior of a structure.
Floating debris is assumed to travel at the flow
velocity, as confirmed by laboratory experiments on a Debris impact loads need
one-fifth scale shipping container in the Large Wave not be considered to act
Flume at Oregon State University (Riggs et al. 2014). simultaneously on multiple
Debris impact is considered to be a single concentrated members.
force acting anywhere within the flow depth since the The impact duration is
debris may not always travel at the water surface. The extremely short, and the
impact loads must be applied to the structural member probability of multiple
simultaneous strikes is
at locations that cause maximum bending moment and
therefore very low.
maximum shear force in the member.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Downloaded from ascelibrary.org by La Trobe University on 10/20/20. Copyright ASCE. For personal use only; all rights reserved.

Debris Impact Loads

Figure G11-1. Flowchart for debris impact.

Tsunami Loads and Effects


@Seismicisolation
@Seismicisolation
159
160 Tsunami Loads and Effects

Debris impact loads need not be considered to act simultaneously on multiple mem-
bers since the impact duration is extremely short and the probability of multiple simultane-
ous strikes is very low. In addition, other tsunami-related loads, such as hydrodynamic and
hydrostatic loads, need not be combined with the debris impact loads. Because of conserva-
tism in the debris impact loading and the low probability that a direct strike from a shipping
container, log, or other large debris object occurs at the peak hydrodynamic or hydrostatic
load, it is not required that these loads be considered simultaneously.
Debris strikes are highly impulsive in nature and will result in dynamic excitation of the
Downloaded from ascelibrary.org by La Trobe University on 10/20/20. Copyright ASCE. For personal use only; all rights reserved.

impacted structural element. When debris strikes a structural element, the impact dura-
tion can be extremely short. If this duration is similar to the natural period of vibration of
the structural element, the dynamic effects can lead to an increased force in the element.
This dynamic amplification can be approximated by using the Rmax factor from ASCE 7-16,
Table  6.11-1. However, it is more accurate, and generally less conservative, to consider
dynamic analysis of the structural element response to a rectangular pulse as described in
ASCE 7-16, Section 6.11.8. This approach can be used for both elastic and inelastic structural
response and will generally produce the most accurate structural design. For impacts in
which the structural element exceeds its elastic limit, work-energy methods may also be
used (ASCE 7-16, Section 6.11.8).

11.1  Alternative Simplified Debris Impact


Static Load
As the tsunami flow velocity increases, so do the potential impact forces from debris floating
in that flow. However, if the impact force exceeds the axial capacity of the debris material,
then the force will be limited to the material axial capacity. For a log impact, unless the end
of the log is perfectly flat and strikes the structural element perfectly perpendicular, it is
appropriate to assume that only a portion of the log end area will be crushed during the
impact. ASCE 7-16 assumes a contact area equal to 20% of the cross-sectional area of a 1 ft
diameter log, or approximately 22 in.2 Assuming a timber crushing strength of 5,000 psi, rep-
resenting approximately the mean plus one standard deviation for Southern pine or Doug-
las fir according to ASTM D2555 (2017), the resulting crushing force is 110 kips. Including a
dynamic response factor of 1.5 results in the basic direct strike force of Fi = 165CoITSU kips.
This represents 50% of the force given in ASCE 7-16, Section 6.11.1, which applies to sites that
are not in an impact zone for shipping containers, ships, and barges.
Similarly, tests at Lehigh University using a full-scale 20 ft shipping container showed
that the maximum impact force was limited to the combined yielding and buckling strength
of the axial member running along the bottom corners of the container (Piran Aghl et al.
2014). These are the strongest structural elements making up the container shell, and their
axial stiffness results in the largest impact loads if the container is oriented with its long axis
parallel to the flow. For this condition, the maximum impact force was 311 kips when both

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Debris Impact Loads 161

bottom corners were struck simultaneously (Piran Aghl et al. 2014). Because the ASCE 7-16
provisions assume a strike at only one bottom corner and the force is proportional to
k � , for half the stiffness the impact force at one corner is 311 / 2 = 220 kips. Including a
dynamic amplification factor of 1.5 results in the prescribed static load of 330 kips. Trans-
verse impacts were also considered, but the maximum impact force was less than that for
an axial strike.
When contents were added to the container, there was no increase in the magnitude of
the impact force because the impact duration is too short to activate the mass of the con-
Downloaded from ascelibrary.org by La Trobe University on 10/20/20. Copyright ASCE. For personal use only; all rights reserved.

tents unless they are rigidly attached to the container shell. However, the duration of the
resulting impulse increased to absorb the momentum of the contents.

11.2  Wood Poles and Logs


Extensive laboratory testing and theoretical modeling as part of a multiyear Network for
Earthquake Engineering Simulation Research (NEESR) project have led to the development
of an improved understanding of the impact loads resulting from floating debris (Riggs et al.
2014). In contrast to prior debris impact expressions that were based on the impulse-mo-
mentum theory, this new approach to impact loads is based on the stress wave propaga-
tion along the length of the member. The assumption is that the debris element, such as a
log or shipping container, is oriented longitudinally in the direction of flow as it strikes the
structural element. An elastic compressive stress wave propagates along the length of the
member and is reflected as a tensile stress wave which returns to the impact end, leading to
separation of the debris from the structural element.
The duration of impact is therefore based on the wave propagation velocity in the mate-
rial, c0  E /  , where E and ρ are the elastic modulus and mass density of the debris mate-
rial, respectively. If the length of the debris is L, then the time taken for the stress wave to
travel along the debris length and return to the impact end is td = 2L/c0. Riggs et al. (2014)
show that the analytical solution to the one-dimensional (1D) wave equation results in an
impact force given by F  E  Aumax , where A is the debris cross-sectional area and umax is
the debris velocity immediately before impact. For a purely elastic impact, the impulse can
be represented by a rectangular pulse with magnitude F and duration td. In reality, plastic
deformation at the impact point results in some distortion of the pure rectangular pulse, but
as shown in Figure G11-2 from Piran Aghl et al. (2014), both wooden logs and shipping con-
tainers produce load pulses that closely match the analytical solution.
Given that the total mass of the debris is md = ρAL and the axial stiffness of the debris is
k = EA/L, the impact force can also be written as Fni = umax kmd , which is given in ASCE 7-16
as Equation (6.11-2). This equation has the same form as Equation (6-8) in FEMA P-646, second
edition, which gives the floating debris impact force as Fi  1.3umax kmd (1  c) , where the 1.3
factor is an importance coefficient for Risk Category IV structures, and c is a hydrodynamic
mass coefficient which represents the effect of fluid in motion with the debris (FEMA 2012).

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


162 Tsunami Loads and Effects
Downloaded from ascelibrary.org by La Trobe University on 10/20/20. Copyright ASCE. For personal use only; all rights reserved.

Figure G11-2. Nondimensionalized load time histories: (a) shipping container with two-bottom-corner
strike; (b) wood pole large end strike.
Source: Piran Aghl et al. (2014).

Given that the theoretical impact duration is


2L 2L
td  
c0 E

and the impact force is Fni = E r Aumax , a little manipulation shows that

Fni E

umax A 
and
2umax  AL 2umax md
td  
Fni Fni
which is the expression given by ASCE 7-16, Equation (6.11-4).
@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Debris Impact Loads 163

These equations are based on in-air analytical solution and laboratory tests. The effect
of performing the same impact in water was studied at Oregon State University by Ko (2013).
It was found that for a one-fifth scale aluminum shipping container model, the impact force
in water was typically within 10% of the force predicted by the in-air analytical solution
(Figure G11-3). It was concluded that for design purposes, the conservatism in assuming a
direct longitudinal impact at a velocity equal to the “free-field” water velocity “is likely to be
sufficient to cover the unconservatism in not including the fluid effect” (Riggs et al. 2014).
The effect of the surrounding fluid is ignored in the ASCE design approach.
Downloaded from ascelibrary.org by La Trobe University on 10/20/20. Copyright ASCE. For personal use only; all rights reserved.

Debris impact on a structural element is a dynamic event, characterized by a large-


magnitude force acting over a very short duration. ASCE  7-16 permits the design of the
structural element for an equivalent static load, which may or may not be conservative. If the
impact duration is greater than about 25% of the natural period of the impacted structural
element, then the dynamic response of the structural element will exceed the applied static
force. For this reason, the static load must be multiplied by a dynamic response factor, Rmax,
given in ASCE 7-16, Table 6.11-1. The peak Rmax of 1.8 occurs for an impact duration of 0.7 to
0.9 times the structural period, implying an 80% magnification effect owing to the dynamics
of the structural element. The Rmax values provided in ASCE 7-16, Table 6.11-1 are based on
ASCE 7-16, Table C5.4-4.
In lieu of the static analysis incorporating the response factor Rmax, it is also permissi-
ble to perform a full dynamic analysis of the structural element subjected to a rectangular
impulse with magnitude given by Equation (6.11-2) and duration given by Equation (6.11-4).


Figure G11-3. Comparison of in-air and in-water impact tests using one-fifth scale aluminum shipping
container model at Oregon State University: (left) comparison of individual tests; (right) comparison
of all in-water tests with different flow depths, velocities, and container contents against the straight
line representing the analytical in-air solution.
Source: Ko (2013); (Right) adapted from Riggs et al. (2014).

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


164 Tsunami Loads and Effects

The preceding analytical and experimental studies, on which ASCE 7-16, Equation (6.11-2)


is based, were all performed assuming that the debris strike is perfectly aligned with the
longitudinal axis of the stiffest structural element of the debris. This represents the worst
condition but has a relatively low likelihood of occurring during an actual tsunami. Haehnel
and Daly (2004) performed debris impact experiments using wooden logs initially oriented
longitudinally in the flow. No attempt was made to control the orientation of the log as it
was carried by the flow toward the target column. As a result, many of the logs completely
missed the target and others resulted in glancing blows in addition to a few direct strikes.
Downloaded from ascelibrary.org by La Trobe University on 10/20/20. Copyright ASCE. For personal use only; all rights reserved.

This was not a completely random study of debris orientation because of the initial align-
ment of the log with the flow, but it provides a measure of the probability of a direct strike.
Considering only the logs that impacted the target, and based on the frequency distribution
of impact angle, the mean impact force was 33.5% of the direct impact force, with a stan-
dard deviation of 32.4%. To maintain a similar reliability for debris impact strikes as was
used for the momentum flux determined by the energy grade line analysis (see Section 6.2),
one standard deviation was added to the mean to get an orientation factor of approximately
65%, or Co = 0.65. Clearly, if the logs had not been oriented parallel to the flow, and if all the
misses were included in the statistical analysis, the probability of a direct strike would be
reduced, and the orientation factor would be lower. Because of this inherent conservatism
in the selection of Co = 0.65, it was considered reasonable to ignore any hydrodynamic load
that might occur simultaneously with a debris strike. Example G11-1 demonstrates the cal-
culation of debris impact due to a log.

11.3  Debris Impacts by Floating Vehicles


Past tsunamis and other flooding events provide ample evidence that passenger vehicles will
float and be transported by the flow until water is able to seep into the passenger compart-
ment, which might take many minutes. It is therefore required that all structures located
where the tsunami flow depth exceeds 3 ft consider impact caused by floating vehicles. Mod-
ern passenger vehicles are designed with crumple zones to reduce the impact force during
accidents and extend the impact duration. Based on a National Crash Analysis Center (NCAC)
study of a 2,400 lb vehicle impact (NCAC, “Extended validation of the finite element model
for the 2010 Toyota Yaris passenger sedan,” working paper, 2012), the initial stiffness of a 35
mph collision was estimated to be 5,700 lb/in. (Naito et al. 2014). With the vehicle floating in
a 9 mph (13 ft/s) tsunami flow, Equation (6.11-2) gives

Fni = umax kmd = 13 5, 700 ×12 × 2, 400 / 32.2 = 29, 350 lb ≈ 30 kips

ASCE 7-16 therefore requires that an impact force of Fni = 30 kips be applied to all exte-
rior structural elements from 3 ft elevation to the maximum inundation depth.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Debris Impact Loads 165

Example G11-1. Log Debris Impact Example.

Consider an exterior column of the prototypical MRF office building introduced in


Chapter 2. Assume a flow depth of 16 ft and a flow velocity of 18 ft/s.
The nominal maximum instantaneous debris impact force is given by

ni F =u d km
max
(6.11-2)
Downloaded from ascelibrary.org by La Trobe University on 10/20/20. Copyright ASCE. For personal use only; all rights reserved.

where
umax = 18.0 ft/s (5.5 m/s),
k = EA/L for the wood log with a minimum value of 350 k/in. = 4.2 × 106 lb/ft
(61,300 kN/m), and
md = 1,000/32.2 = 31.1 sl for the minimum 1,000 lb log (454 kg).

Therefore,

Fni = umax kmd = 18.0 4.2 ×106 × 31.1 / 1, 000 = 206 kips

[ Fni = umax kmd = 5.5 61, 300, 000 × 454 / 1, 000 = 917 kN]

For a TRC II building, the importance factor is ITSU = 1.0. The design instantaneous
debris impact force is then given by

Fi = ITSU Co Fni = 1.0 × 0.65× 206 = 134 kips (596 kN) (6.11-3)

The impulse duration is given by ASCE 7-16, Equation (6.11-4) as

2md umax 2  31.1 18.0


td    0.0054 s

Fni 206, 000
2md umax 2  454  5.5
[td    0.0054 s]

Fni 917, 000

The column can be designed using a dynamic analysis by applying an impulsive


rectangular pulse with magnitude Fi and duration td. Alternatively, an equivalent elastic
static analysis can be performed of the column subjected to Fi multiplied by a dynamic
response factor, Rmax, given in ASCE 7-16, Table 6.11-1. The ratio of impact duration to nat-
ural period of the impacted structural element is obtained using td and the natural period
of the column assumed to be fixed-fixed. For this case, the natural period is given by

 L2  
Tcol  2  
 22.373  EI

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


166 Tsunami Loads and Effects

where
L = Unbraced column length = 14 ft – 22 in./12 = 12.17 ft (3.71 m),
ρ = Column mass per unit length = 2.5 ft × 2.5 ft × 150 pcf/32.2 ft/s2 = 29.1 sl/ft
[ = 0.76 × 0.76 × 2,400 kg/m3 = 1,386 kg/m],
E = Modulus of elasticity of the column concrete = 3,600 ksi = 518.4 × 106 psf
(24.9 GPa), and
I = Moment of inertia of column section = bd3/12 = 2.5 × 2.53/12 = 3.26 ft4 (0.0278 m4).
Downloaded from ascelibrary.org by La Trobe University on 10/20/20. Copyright ASCE. For personal use only; all rights reserved.

Therefore,

 L2    12.17 2  29.1
Tcol  2    2    0.00546 s
 22.373  EI  22.373  518.4  10  3.26
6

 L2    3.712  1386
[Tcol  2    2    0.00546 s]
 22.373  EI  22.373  24.9 10  0.0278
9

The ratio of impact duration to column natural period is therefore

td/Tcol = 0.0054/0.00546 = 0.989

Table 6.11-1 gives the dynamic response factor Rmax = 1.7; therefore, the equivalent
static load is given by

Fes = RmaxFi = 1.7 × 134 = 228 kips ( =1.7 × 596 = 1,013 kN)

This exceeds the maximum required impact force from the alternative simplified
impact load given by Section 6.11.1 as 50% of Fi from ASCE 7-16, Equation (6.11-1)

Fi = 330CoITSU = 330 × 0.65 × 1.0 = 215 kips

[Fi = 1,470CoITSU = 1,470 × 0.65 × 1.0 = 956 kN]

So for log impact, Fi = 0.5 × 215 = 107.5 kips (478 kN)


Therefore, the column can be evaluated for a lateral static point load of 107.5 kips
(478 kN) applied at locations that are critical for flexure and shear.

11.4  Debris Impacts of Submerged Tumbling


Boulder and Concrete Debris
Tsunami flows have been observed to move large boulders and pieces of concrete rubble
(Figure G11-4). These objects are assumed to roll along the ground surface and impact any
exterior structural elements at 2 ft above grade. Because of their mass, a minimum flow
depth and velocity are required to initiate and maintain boulder movement. ASCE 7-16 only
requires that boulder impact be considered when the flow depth exceeds 6 ft. The impact

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Debris Impact Loads 167

force is based on a simplified static approach (Chau and Bao 2010) considering a 5,000 lb
boulder tumbling at a velocity of 13.1 ft/s, with a dynamic amplification factor of 2. The result-
ing force of 8,000 lb must be applied at 2 ft above grade on all exterior structural elements.
If the ground-floor slab-on-grade is designed as an isolated slab to avoid buoyancy effects,
it might also be necessary to consider tumbling debris impact on interior columns, although
not specifically required by the ASCE 7-16 standard.
Downloaded from ascelibrary.org by La Trobe University on 10/20/20. Copyright ASCE. For personal use only; all rights reserved.

Figure G11-4. Rolling debris impacts: (left) in Taro, Japan (N 39.7362; E 141.9752); and (right) in Malaela,
Samoa (S 14.0204; W 171.4250).
Source: (a) Courtesy of David Kriebel.

11.5  Site Hazard Assessment for Shipping


Containers, Ships, and Barges
Shipping containers, ships, and barges present a particularly severe form of floating debris
impact loading. Although shipping containers can be found in various locations throughout
coastal communities, they are frequently concentrated at container yards associated with
ports, railways, or other transportation hubs. ASCE 7-16 does not require that all structures
consider the potential impact from shipping containers because of the relatively remote
possibility of this occurring unless the site is located near a shipping container storage yard.
Similarly, impact from ships and barges need only be considered by structures close to a port
or harbor.
Naito et al. (2014) used satellite images of the post-Tohoku tsunami debris distribution
to develop a zone of influence for ships, boats, and shipping containers. This methodol-
ogy was further developed by the TLESC for incorporation into the ASCE  7-16, Chapter 6
provisions. For every concentrated source of shipping containers, ships, or barges, a debris
impact hazard region must be determined. The procedure assumes that all the debris objects
initiate from the centroid of the storage yard, in the case of shipping containers, or from the
center of the port, in the case of ships and barges. If the storage yard or port are distributed

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


168 Tsunami Loads and Effects

longitudinally along the coast, it may be necessary to represent the debris source as a num-
ber of individual point sources.
The procedure to develop the debris impact hazard region is as follows:

1. Determine the plan area of all debris objects at the source based on the average number
of containers, barges, or ships, multiplied by their plan area.
2. Determine the geometric center of the debris source.
3. Determine the primary flow direction perpendicular to the shoreline and passing
Downloaded from ascelibrary.org by La Trobe University on 10/20/20. Copyright ASCE. For personal use only; all rights reserved.

through the debris source centroid.


4. Project lines rotated clockwise and counterclockwise about the debris source centroid
by 22.5 degrees from the primary flow direction.
5. This will develop a 45 degree wedge extending inland from the debris source centroid.
6. This wedge terminates with a circular arc segment at a radius such that the area of
the resulting circular sector is 50 times the total sum area of the debris at the source.
This implies that the debris is distributed over an area resulting in 2% concentration,
compared with the 100% concentration assumed at the source. This accounts for the
distribution of debris objects during inflow.
7. If the circular sector extends beyond the inundation limit shown in the tsunami design
zone map, then clearly the debris objects will run aground and not proceed. ASCE 7-16
assumes that this grounding will occur at an inundation depth of 3 ft for shipping con-
tainers, or the ballasted draft plus 2 ft in the case of barges and ships.
8. The extent of the circular sector may also be curtailed if structural steel and/or con-
crete structures of sufficient height block the debris (Figures G11-5 and G11-6).
9. During outflow, the debris objects will be drawn back toward the coastline. A “flipped” distri-
bution wedge is drawn from the intersection of the primary flow direction and the circular
arc segment, or curtailed debris flow extent, and extending back to the coastline. Again, the
wedge is formed by lines oriented 22.5 degrees either side of the primary flow direction.
10. Structures located only in the inflow circular sector need only consider debris impact on
their coastal elevation, whereas structures located only in the outflow sector need only
consider large debris impact on their inland elevation. Structures located in both inflow and
outflow sectors must consider large debris impacts on both coastal and inland elevations.

11.6  Shipping Containers


The impact force for shipping container impacts is determined using ASCE 7-16, Equations
(6.11-2) and (6.11-3), where md is the mass of the empty container. The mass of contents need
not be included in calculation of the impact force but must be considered when computing the
impact duration. Full-scale shipping container impact tests performed at Lehigh University
showed that the mass of contents does not affect the impact force significantly unless they
are rigidly attached to the structural frame of the container, which is seldom the case (Piran

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Debris Impact Loads 169

Aghl et al. 2014). However, the same experiments indicated that the presence of contents may
increase the duration of impact, so one is required to consider a range of impact durations
from empty to fully loaded containers. For an empty container, the minimum debris impact
duration is given by ASCE 7-16, Equation (6.11-4). For a fully loaded container, the maximum
debris impact duration is given by ASCE 7-16, Equation (6.11-5). Because the actual container
contents are unknown, one must consider the worst impact duration within this range.
Downloaded from ascelibrary.org by La Trobe University on 10/20/20. Copyright ASCE. For personal use only; all rights reserved.

Figure G11-5. Google Earth images of Sendai Port shipping container yard before (top) and after
(bottom) the Tohoku tsunami. Note how steel and concrete framed buildings to the west of the
container yard have blocked the containers from progressing further inland (N 38.2659; E 141.0223).
Source: Google Earth, images (top) August 31, 2008, (bottom) March 23, 2011.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


170 Tsunami Loads and Effects
Downloaded from ascelibrary.org by La Trobe University on 10/20/20. Copyright ASCE. For personal use only; all rights reserved.

Figure G11-6. Shipping containers at Talcahuano Port in Chile are prevented from moving further
inland by reinforced concrete and structural steel buildings adjacent to the port during the 2010
tsunami (S 36.715; W 73.109).
Source: Courtesy of IFRC.org, International Federation of Red Cross and Red Crescent Societies.

Example G11-2. Site Hazard Assessment.

Naito et al. (2016) demonstrate the application of the debris impact hazard assess-
ment tool at a shipping container storage facility adjacent to Hilo Harbor, Hawai‘i (Fig-
ure G11-7). The authors assumed a worst case of 514 20 ft (6.1 m) and 1,418 40 ft (12.2 m)
standard shipping containers at the container yard. This example is used to demon-
strate the application of the debris impact hazard assessment tool as follows:
Step 1. Determine the plan area of all debris objects at the source based on the
number of containers multiplied by their plan area. A standard shipping container is
8 ft (2.43 m) wide; therefore, the total plan area of the 20 ft and 40 ft containers is
A = 514 × (20 × 8) + 1,418 × (40 × 8) = 536,000 ft2 (49,657 m2)
Step 2. Determine the geometric center of the debris source. Naito et al. (2016)
identify the center of the storage yard as shown in Figure G11-7.
Step 3. Determine the primary flow direction perpendicular to the shoreline and
passing through the debris source centroid. The flow transect is indicated by the red
line in Figure G11-7.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Debris Impact Loads 171

Step 4. Project lines rotated clockwise and counterclockwise about the debris
source centroid by 22.5 degrees from the primary flow direction (Figure G11-7).
Steps 5 and 6. The resulting 45 degree wedge extends inland until the circular arc
segment is 50 times the total area of the debris at the source. Therefore, the area of the
arc segment must equal 50A = 26,800,000 ft2 (2,482,850 m2). The area of a 45 degree arc
segment is one-eighth of the full circle area; therefore, the radius of the arc segment
is determined as follows:
Downloaded from ascelibrary.org by La Trobe University on 10/20/20. Copyright ASCE. For personal use only; all rights reserved.

 R2
Aarc   50 A  26, 800, 000
8

Therefore, R = 8,261 ft (2,514 m). This arc is drawn in blue in Figure G11-7.
Step 7. If the circular sector extends beyond the inundation limit, then the debris
is assumed to run aground at an inundation depth of 3 ft. As shown in Figure G11-7, this
occurs at the Hilo site where the 3 ft inundation depth line is shown in red.
Step 8. The circular sector would also be curtailed if structural steel and/or con-
crete structures block the debris, but no such structure exist at the Hilo location.
Step 9. A “flipped” distribution wedge is now drawn from the intersection of the
primary flow direction and the curtailed debris flow extent as shown in Figure G11-7.
The total shaded area in Figure G11-7 represents the region exposed to potential ship-
ping container strikes.
Step 10. Any building or structure requiring tsunami design and located in this
shaded area must consider shipping container impact to the exterior structural mem-
bers. In the center diamond-shaped region, these impacts could occur on both ocean
and inland faces of the building. In the two inland wedges, ASCE  7-16 only requires
that strikes be considered on the ocean-facing side of the building. In the two coastal
wedges, ASCE 7-16 only requires that impact be considered on the inland side of the
building; however, it may be prudent to consider both ocean and inland faces because
of the potential for debris strikes during subsequent waves.

The impact force for shipping containers is based on the worst case assumption of a
direct axial strike between one end of a bottom corner member of the container and the
structural element being considered (Figure G11-8). Minimum values for the weight and
stiffness of the containers are given in ASCE 7-16, Table 6.11-2, but some containers might
exceed these values. The actual properties of typical containers at the storage facility under
investigation should be used if they exceed the values in Table 6.11-2.
The velocity of the shipping container is assumed to match the flow velocity corre-
sponding to the flow depth being considered. Because the shipping container is assumed
to float at the surface of the flow, this would imply that the impact force will vary as the

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


172 Tsunami Loads and Effects
Downloaded from ascelibrary.org by La Trobe University on 10/20/20. Copyright ASCE. For personal use only; all rights reserved.

Figure G11-7. Application of debris impact hazard assessment tool at Hilo Harbor, Hawai‘i.
Source: Naito et al. (2016); Courtesy of ASCE Journal of Waterway, Port, Coastal, and Ocean Engineering.

flow depth increases. In lieu of considering this variation, the design can be based con-
servatively on the maximum flow velocity determined at the site. Once the impact force
exceeds the elastic limit of the axial corner member of the shipping container, this member
will yield and/or buckle, thus limiting the maximum impact force. Based on the experiments
at Lehigh University, this maximum impact force for a standard 20 ft long shipping container
was 330,000 lb. Regardless of the flow velocity, one need not consider a shipping container
impact force exceeding this limit.
As discussed for log impacts, the potential for a direct axial strike on the bottom cor-
ner of a shipping container is relatively low. Lacking any experimental or simulation data to

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Debris Impact Loads 173
Downloaded from ascelibrary.org by La Trobe University on 10/20/20. Copyright ASCE. For personal use only; all rights reserved.

Figure G11-8. Standard 20 ft shipping container showing assumed direct strike along bottom beam
element.

justify an appropriate orientation factor to adjust for this probability, the ASCE 7-16 standard
uses the same Co = 0.65 determined from the Haehnel and Daly (2004) study on log impacts.

11.7  Extraordinary Debris Impacts


Buildings and other structures located within or close to ports and harbors are exposed to
potential impact from boats, barges, and ships (Figure G11-9). When the building or structure
is located within the debris hazard region of a pier or wharf where barges or ships may berth,
ASCE 7-16 requires that extraordinary debris impacts be considered. This is only required for
TRC III critical facilities and TRC IV buildings and structures, and only if the maximum inun-
dation depth at the site exceeds 12 ft, because of the larger draft of these vessels. The impact
force is determined using ASCE 7-16, Equation (6.11-2), using a debris mass determined from
the lightweight tonnage (LWT) plus 30% of the deadweight tonnage (DWT), and the stiffness
of the impacted structural element. The resulting force must be applied at all critical loca-
tions along the height of the structural member from the base of the building up to 1.3 times
the inundation depth plus the height of the vessel deck above the waterline (Figure G11-10).
Because of the potentially large magnitude of these extraordinary debris impact forces,
it may not be feasible or economical to design the individual structural elements to resist
these loads (Figure G11-11). ASCE 7-16 allows for the incorporation of alternative load path
progressive collapse prevention following ASCE 7-16, Section 6.8.3.5.3 in lieu of strengthening
the individual members. The alternative load path provisions must be applied to all structural

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


174 Tsunami Loads and Effects
Downloaded from ascelibrary.org by La Trobe University on 10/20/20. Copyright ASCE. For personal use only; all rights reserved.


Figure G11-9. Fishing vessel impacts with (left) building, and (right) bridge pier, in Miyako City, Japan,
during the Tohoku tsunami.

members from the base of the structure up to the story level above 1.3 times the inundation
depth plus the height of the ship deck measured from the waterline (Figure G11-10).

11.8  Alternative Methods of Response Analysis


In lieu of the simplified static approach described previously, structural members can be
designed for impact loads using a dynamic analysis for both elastic or inelastic impacts, or a
work-energy method for inelastic impacts. The forcing function is represented by a rectan-
gular pulse with magnitude given by ASCE 7-16, Equation (6.11-2) and duration td. The debris
velocity used in the work-energy method of analysis must be the maximum velocity, umax,
multiplied by the importance factor, ITSU, and the orientation factor, Co, to determine the
appropriate kinetic energy before impact.

11.9  Detailed Debris Impact Calculation for


Monterey, California, Office Building
The 6-story MRF office building located near the shoreline in Monterey, California (see
Chapter 2) is not within the impact hazard zone of a port, harbor, or shipping container
storage facility. Therefore, impacts from shipping containers, boats, and ships need not be
considered. However, impact from wood logs and poles, vehicles, and tumbling boulders and
concrete debris must be considered as demonstrated in this section. The controlling flow
depth and velocity at the project site are 21.5 ft and 31.5 ft/s, respectively, as determined
from EGLA.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Debris Impact Loads 175
Downloaded from ascelibrary.org by La Trobe University on 10/20/20. Copyright ASCE. For personal use only; all rights reserved.

Figure G11-10. Illustration of barge or ship impact provisions.

11.9.1  Wood Logs and Poles: Section 6.11.2


The calculation of log debris impact force on an exterior column of the MRF office building
is similar to Example G11-1 and again results in the maximum impact load of Fi = 107.5 kips
(478 kN) governed by the crushing strength of the log.

11.9.2  Impact by Vehicles


ASCE  7-16, Section 6.11.3 provides the impact force resulting from floating vehicles. The
impact force is given as

Fi = ITSU × 30 = 1.0 × 30 = 30 kips

This will not control over the log impact load.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


176 Tsunami Loads and Effects
Downloaded from ascelibrary.org by La Trobe University on 10/20/20. Copyright ASCE. For personal use only; all rights reserved.

Figure G11-11. Cargo ship impact with warehouse and crane in Sendai Harbor during the Tohoku
tsunami. Note the deck-level strike with the steel column supporting the roof and the submerged bulb
impact with the concrete foundation wall (N 38.26865; E 141.01055).

11.9.3  Impact by Submerged Tumbling Boulder


and Concrete Debris
ASCE 7-16, Section 6.11.4 gives the impact force that must be considered for tumbling boul-
ders and concrete debris. These impacts need only be considered when the flow depth
exceeds 6 ft. Because hmax = 21.5 ft > 6 ft, an impact force of Fi = ITSU × 8,000 = 8,000 lb shall
be applied at 2 ft above grade. This will not control over the log impact load.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


12
Foundation Design
Downloaded from ascelibrary.org by La Trobe University on 10/20/20. Copyright ASCE. For personal use only; all rights reserved.

Many structural failures during past tsunamis are the result of scour around shallow foun-
dations (Figure G12-1). This chapter describes the provisions for foundation design of both
deep and shallow foundations subjected to both general erosion and local scour. Prescriptive
design provisions are provided, along with the option to use performance-based criteria.
Remedial measures are also provided to mitigate general erosion and local scour. Design
provisions also cover the effects of plunging scour as flow passes over a rigid object (Fig-
ures G12-2 and G12-3).

Figure G12-1. Four foot (1.4 m) deep scour pit below building foundation in Taro, Japan (N 39.7361;
W 141.9754).

177
@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


178 Tsunami Loads and Effects
Downloaded from ascelibrary.org by La Trobe University on 10/20/20. Copyright ASCE. For personal use only; all rights reserved.

Figure G12-2. Plunging scour behind flood barrier in Miyako, Japan, caused by overtopping (N 39.6399;
E 141.9553).

Figure G12-3. Plunging scour behind tsunami barrier in Miyako, Japan (N 39.5871; E 141.9499).

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Foundation Design 179

Typical foundation failure mechanisms are

1. Lateral sliding attributable to tsunami hydrostatic and hydrodynamic forces, with the
added effects of any unbalanced lateral soil pressure caused by unequal soil levels on
either side of the structure or individual foundation, exacerbated by the effects of local
scour on only one side of the foundation.

2. Uplift or flotation caused by buoyancy effects acting on the overall structure and the
Downloaded from ascelibrary.org by La Trobe University on 10/20/20. Copyright ASCE. For personal use only; all rights reserved.

foundations (Figure G9-1). Pile tensile capacity required to counteract potential uplift


must be determined by taking into account potential local scour around the top of the
pile and loss of soil shear strength caused by pore pressure softening and prior lique-
faction if the tsunami is the result of a near-source earthquake.

3. Piping conditions that can develop owing to excess seepage stresses reducing the
strength and integrity of the soil around and below the foundation.

4. Slope stability failure and lateral spreading caused by soil saturation and pore pressure
softening resulting from tsunami inundation and drawdown.

5. Bearing capacity failure, particularly where sustained pore pressure because of tsu-
nami inundation reduces the soil strength properties. Because it is not expected to be a
common failure mechanism, bearing capacity need only be checked where specifically
required in the ASCE 7-16 provisions.

Figure G12-4 shows extensive landside scour at a seawall intended to protect the Sendai,
Japan, airport. The scour is likely the result of both plunging and overtopping scour and from
longitudinal diversion during drawdown, collecting at intermittent breach points approxi-
mately every 100 m along the wall. Granular fill materials supporting the wall were readily
eroded. At the same location, Figure G12-5 shows similar plunging scour behind a buttressed
cantilever reinforced concrete wall that survived the tsunami. Figure G12-6 shows scour
around the corner of a building foundation in Malaela, Samoa, and Figure G12-7 shows scour
around a bridge pier in Miyako, Japan.

12.1  Resistance Factors for Foundation


Stability Analyses
The maximum considered tsunami is specified as the design condition for both structural
design of the building and geotechnical design of the foundations and any remedial measures
being considered. As with other geotechnical design, soil loading analysis during a tsunami
requires geotechnical judgment in selecting a reduced nominal strength that adequately

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


180 Tsunami Loads and Effects
Downloaded from ascelibrary.org by La Trobe University on 10/20/20. Copyright ASCE. For personal use only; all rights reserved.

Figure G12-4. Scour behind seawall protecting Sendai, Japan, airport (N 38.1336; E 140.9414).
Source: Courtesy of Mathew Francis.

Figure G12-5. Undamaged buttressed structural seawall protecting Sendai, Japan, airport with
significant sand accumulation on ocean side (left) and up to 13 ft (4 m) deep scour pockets on land side
(right) (N 38.1321; E 140.9408).
Source: Courtesy of Mathew Francis.

represents the inherently nonlinear behavior of soil materials. To ensure that the assumed
failure does not occur, the loading condition must be less than the material nominal strength
multiplied by a resistance factor. ASCE 7-16 Commentary, Section C6.12.1 gives background
to the selection of 0.67 (equivalent to a “safety factor” of 1.5) as the resistance factor for geo-
technical design when applied to tsunami conditions.

12.2  Load and Effect Characterization


Foundations and tsunami barriers are required to be designed for all the loads and effects
presented in ASCE 7-16, Sections 6.12.2.1 through 6.12.2.6. In addition, the effect of soil
strength loss, general erosion, and scour must also be considered following the requirements
given in this section. Because a tsunami commonly consists of more than a single wave, it is

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Foundation Design 181
Downloaded from ascelibrary.org by La Trobe University on 10/20/20. Copyright ASCE. For personal use only; all rights reserved.

Figure G12-6. Scour below corner of building footing in Malaela, Samoa (S 14.0264; W 171.4293).

Figure G12-7. Local scour of clayey soil around bridge piers in Miyako, Japan (N 39.6410; E 141.9561).
Source: Courtesy of Mathew Francis.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


182 Tsunami Loads and Effects

important that the effect of subsequent waves is considered. Laboratory experiments have
shown that general erosion and scour are more pronounced during the drawdown phase of
tsunami inundation than owing to the incoming waves (Yeh and Li 2008, Xiao et al. 2010).
This is attributed to the effect of pore pressure softening when the drawdown occurs more
rapidly than the release of pore pressure in the soil. The result is rapid scour of a “heavy liq-
uid” of water with large sediment content. It is therefore anticipated that significant scour
will occur after withdrawal of the first major inundation. Subsequent waves will then load
the structure when the foundations may have been significantly compromised by this gen-
Downloaded from ascelibrary.org by La Trobe University on 10/20/20. Copyright ASCE. For personal use only; all rights reserved.

eral erosion and scour. It is therefore required that the designer consider a minimum of two
wave cycles, the first having an inundation depth of 0.8hmax, and the second having an inun-
dation depth of hmax for the particular location (ASCE 7-16, Section 6.8.8). The hydrostatic
and hydrodynamic loading from the second wave must therefore be applied to the structure
after it has already experienced general erosion and scour resulting from the first wave.
ASCE 7-16 Commentary, Figure C6.12-1 gives a schematic of the various tsunami loading
conditions for a typical foundation element (Figure G12-8). The effects of general erosion are
assumed to be equal over the entire project site, whereas local scour is likely to vary around
the overall foundation and around individual footings, depending on their size. For piles that
are exposed to scour it is adequate to assume that the scour depth is equal on all sides of the
pile; however, for large spread footings there is a potential for differential scour and even
undermining of the footing depending on its depth relative to the anticipated scour depth
at the site (Figure G12-1). The deepest scour will concentrate at corners of the foundation
because of accelerated flow velocities at those locations. The designer must determine the
maximum scour depth following the provisions in ASCE 7-16, Section 6.12.2.4 and consider
the potential for differential scour between opposite sides of the foundation element.
This unequal scour on either side of the overall structure or an individual foundation
element will result in unbalanced lateral earth pressures on opposite sides of the foun-
dation. The resulting unbalanced force must be combined with the fluid forces acting on
the exposed portion of the foundation as well as lateral loads acting on the superstructure
above. Figure G12-9 shows a typical condition during the second tsunami wave on a building
with a basement deeper than 12 ft (3.66 m). Assuming hmax = 20 ft at this building location,
then during the first tsunami wave a flow depth of 0.8hmax = 16 ft must be considered. The
local scour around the basement walls owing to hmax = 16 ft will be 12 ft based on ASCE 7-16,
Table 6.12-1 or Figure 6.12-1. Assuming that this scour only occurs on one side of the building,
then during the subsequent inflow or drawdown the foundation system must be designed
for the unbalanced lateral loads caused by hydrodynamic pressure (because of drag on the
entire building) and lateral earth pressure from saturated soil conditions, without the benefit
of passive earth pressure over the scour depth (Figure G12-9).
Figure G12-10 shows the same building but without the basement level. For consolidated
or cohesive soils, the local scour will extend inward and outward from the structure corners
by a distance equal to the scour depth. Exposure of the exterior corner pile foundations will

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Foundation Design 183
Downloaded from ascelibrary.org by La Trobe University on 10/20/20. Copyright ASCE. For personal use only; all rights reserved.

Figure G12-8. Schematic of tsunami loading condition for a foundation element.


Source: ASCE 7-16 Commentary, Figure C6.12-1.

Figure G12-9. Building with basement showing effect of unequal scour on lateral load design of
foundation.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


184 Tsunami Loads and Effects

significantly reduce their lateral stiffness compared with the side and interior piles. One
could compute the relative stiffness and distribute the lateral load caused by hydrodynamic
drag proportionally, or one could choose to ignore the lateral load capacity of the exterior
corner piles and rely only on the side and interior piles for lateral load resistance. The hydro-
dynamic lateral pressure will now apply to both the building above grade and the exposed
grade beam, pile caps, and piles.
For unconsolidated or noncohesive soils, the lateral extent of local scour must be taken
as three times the scour depth. This condition is shown in Figure G12-11. For the given build-
Downloaded from ascelibrary.org by La Trobe University on 10/20/20. Copyright ASCE. For personal use only; all rights reserved.

ing, this results in scour under the entire ground-floor slab, exposing the top of all pile foun-
dations. The lateral load caused by hydrodynamic drag must now be supported by the piles
without assistance of soil over the top 10 to 12 ft of pile. This will require piles with significant
bending and shear capacity.
Because of the potential for pore pressure softening of the saturated upper layers of soil
caused by rapid drawdown, a layer of soil with reduced strength will surround the foundation.
This loss of strength must be accounted for up to a depth of 1.2 times the maximum inunda-
tion depth, hmax, according to ASCE 7-16, Section 6.12.2.2. The strength loss is estimated fol-
lowing the provisions in ASCE 7-16 Commentary, Section C6.12.2.2. Although pore pressure
softening develops from a different mechanism than liquefaction, the resulting decrease in
effective shear strength can significantly affect the foundation design. For near-source tsu-
namis, both effects must be considered because of the short time duration between ground
shaking–induced liquefaction and inundation-induced pore pressure softening.
In addition to the foregoing effects, there is also potential that seepage below the
foundation transfers the increased water pressure from the tsunami inundation into uplift

Figure G12-10. Building with slab-on-grade at ground level showing scour condition for consolidated or
cohesive soils.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Foundation Design 185
Downloaded from ascelibrary.org by La Trobe University on 10/20/20. Copyright ASCE. For personal use only; all rights reserved.

Figure G12-11. Building with slab-on-grade at ground level showing scour condition for unconsolidated
or noncohesive soils.

pressure on the bottom of the foundation. The rate of this increase in pressure below the
footings will depend on the duration of the inundation and the permeability of the soil sur-
rounding the footing. Porous gravels and sandy soils will allow this pore pressure increase
almost coincident with the incoming tsunami flow, so the entire hydrostatic pressure result-
ing from the inundation should be assumed to occur below the foundation. If the soil is a less
porous silty or clayey soil, it may take considerably longer for this pressure to develop, even
if the inundation flow depth is maintained for some time. Section 12.2.1 provides guidance on
estimating this uplift pressure.

12.2.1  Uplift and Under-Seepage Forces


Although traditionally considered only for structures such as dams and levees, hydrostatic
uplift and seepage forces can develop during tsunami inundation owing to the sustained
overland flow or trapping of water in low-lying areas or behind berms, roads, and founda-
tions. Guidance for design in the presence of uplift and seepage is provided by USACE EM
1110-2-2100 (2005) and USACE EM 1110-2-2302 (1990). Careful soil investigations are required
to determine the soil characteristics such as permeability and potential for erosion.
An example of the buoyancy of a large reinforced concrete building in Onagawa, Japan,
during the Tohoku tsunami is shown in Figure G12-12. The incoming flow duration at this
location is estimated from video evidence as approximately 8 min (Koshimura and Hayashi
2012), giving ample time for the pore pressure below the foundation slab to build to the full
hydrostatic pressure, particularly because of the very coarse gravel used as subgrade below
the foundation beams and slabs. Although this building also had piles below the grade beam
system, these piles were purely compression piles. They consisted of hollow concrete pipes

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


186 Tsunami Loads and Effects
Downloaded from ascelibrary.org by La Trobe University on 10/20/20. Copyright ASCE. For personal use only; all rights reserved.

Figure G12-12. Overturned building in Onagawa, Japan, attributed to seepage and uplift pressure below
structural reinforced concrete beam-slab system at grade level. Inset shows spun concrete pipe pile
with inadequate connection to pile cap to resist uplift load (N 38.4407; E 141.4465).

driven into the ground. The connection between the piles and the pile caps was not intended
to resist uplift.
Using this building for illustration, Example G12-1 shows how the geotechnical condi-
tions would be applied assuming a flow depth of 17.5 ft (5.35 m). The estimated flow depth
during the Tohoku tsunami at this location was well over 50 ft (15 m), completely submerging
the 2-story building (Koshimura and Hayashi 2012).
Live and snow loads should not be considered to resist the uplift owing to buoyancy.
In addition, the dead load must be reduced by a factor of 0.9 resulting in the following load
combination:
0.9 D + FTSU < ϕR
where
D = Downward weight of the structure and soil above the bearing surface of the foun-
dation exposed to uplift,
FTSU = Net maximum uplift caused by the distribution of hydrostatic pressure around
the building and foundation for all three load cases defined in ASCE 7-16,
Section 6.8.3.1,
R = Resistance provided by foundation structural elements such as piles and anchors, and
ϕ = Resistance factor of 0.67 for these uplift-resisting elements.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Foundation Design 187

Example G12-1. Pile Foundation Design.

Consider the reinforced concrete refrigerator building in Onagawa, Japan, that floated
and overturned during the Tohoku tsunami. Figure G12-13 shows an isometric view
of the building superstructure with the sealed refrigerator storage room measuring
(12.35 + 9.4) × 8.65 m [71.3 × 28.4 ft] on the ground-floor level.
Downloaded from ascelibrary.org by La Trobe University on 10/20/20. Copyright ASCE. For personal use only; all rights reserved.

Figure G12-13. Dimensions of Onagawa, Japan, refrigerator building.


Source: Chock et al. (2013b).

Figure G12-14 shows a section through the building illustrating the extent of


reduced effective shear stress owing to pore pressure softening. This will affect the
pile tensile capacity to resist uplift and resistance to lateral hydrodynamic loads.
Assuming the maximum inundation depth coincides with the second-floor height
of 17.5 ft, then the uplift pressure caused by seepage below the foundation is equal to

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


188 Tsunami Loads and Effects

the hydrostatic pressure at a depth of 17.5 + 3 = 20.5 ft. This uplift pressure is therefore
1.1 × 64 × 20.5 = 1,443 psf. Applied to a foundation area of 71.3 × 28.4 = 2,025 ft2, the total
uplift force is FTSU = 2,025 × 1,443/1,000 = 2,922 kips (13,000 kN). This uplift force is
greater than the estimated dead weight of the building [2,022 kips (8,995 kN)] (Chock
et al. 2013a).
Applying the load combination, 0.9 D + FTSU ≤ fR, gives –0.9 × 2,022 + 2,922
≤ 0.67 × Rpile
Downloaded from ascelibrary.org by La Trobe University on 10/20/20. Copyright ASCE. For personal use only; all rights reserved.

Therefore, the total pile tension capacity must exceed Rpile = 1,645 kips. This par-
ticular building had 34 piles, meaning that each pile would need to resist an uplift of
1,645/34 = 56 kips per pile. The actual piles consisted of hollow spun concrete pipes
driven into the ground with inadequate connection to the pile caps to resist uplift
as shown in Figure G12-12. These connections were not able to transfer the required
tension, so the building was able to float off the piles. In reality the flow depth at this
location in Onagawa, Japan, exceeded the roof of this building, so without tension piles
it would not be able to resist the uplift pressures.

Figure G12-14. Extent of pore pressure softening and uplift pressure caused by flow inundation.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Foundation Design 189

The moist or saturated unit weight can be used for soil above the saturation level, and
the submerged unit weight must be used for soil below the groundwater table.
A similar foundation failure occurred at a 4-story structural steel framed building in
Onagawa, Japan (Figure G12-15). Although this building had piles below the grade beam sys-
tem, these piles appear to have been designed for compression and shear only. They con-
sisted of hollow concrete pipes driven into the ground. The only connection between the
piles and the pile caps was a short section of spiral reinforcement inserted into the opening
in the pile and grouted over the upper 2 ft of the pile. These “plugs” were probably intended
Downloaded from ascelibrary.org by La Trobe University on 10/20/20. Copyright ASCE. For personal use only; all rights reserved.

to provide shear transfer for seismic loads, but apparently provided insufficient uplift capac-
ity since all failed when the building was subjected to hydrodynamic and/or buoyancy forces.

12.2.2  Loss of Strength


Loss of soil shear strength during tsunami inundation can result from tsunami-induced pore
pressure softening, piping, or seismic shaking caused by the preceding earthquake in sub-
duction zones. Liquefaction owing to seismic shaking is caused by a different mechanism
than pore pressure softening during a tsunami as shown in ASCE 7-16 Commentary, Fig-
ure C6.12-2, but the consequences are the same. Increased pore pressure reduces the effec-
tive shear strength of the soil and increases the potential for sediment transport, scour,
and soil failure. This effect must be assumed to extend to a depth of 1.2 times the maximum
inundation depth at the location. This requirement is based on experiments performed by
Tonkin et al. (2003) on sandy soils in a laboratory flume with relatively short wave periods

Figure G12-15. Overturned 4-story steel framed building in Onagawa, Japan, in which most of the
hollow concrete pipe columns pulled away from the weak pile cap connection, and one pile pulled out
of the ground (N 38.4426; E 141.4463).

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


190 Tsunami Loads and Effects

compared with full-scale tsunami inundation. Given that this depth of pore pressure soften-
ing can occur during relatively short-duration inundation, it is considered likely that similar
depths will be affected even when the soil is less permeable than the sand used in the labo-
ratory experiments. Tsunami-induced pore pressure softening only occurs if the drawdown
is rapid; therefore, this effect need not be considered if the maximum Froude number at the
location is less than 0.5.
Loss of strength owing to pore pressure softening can be evaluated by numerical mod-
eling of soil-structure-fluid interactions, or by multiplying the shear strength by a factor
Downloaded from ascelibrary.org by La Trobe University on 10/20/20. Copyright ASCE. For personal use only; all rights reserved.

1 – Λ, where Λ is the scour enhancement parameter proposed by Tonkin et al. (2003). Exam-
ple G12-2 shows a typical evaluation of the scour enhancement parameter.

Example G12-2. Scour Enhancement Factor, Λ.

ASCE 7 Commentary, Equation (C6.12-2) gives the scour enhancement factor as

 2 hmax  s   2 10  70.4 
  min 0.5,    min 0.5,  
   b cV (Tdraw )  (C6.12-2)
  56 0.5(300) 

where gs is the fluid weight density per ASCE 7-16, Section 6.8.5 = 1.1 × 64 = 70.4 pcf, and
gb is the buoyant weight of the soil (bulk soil density – density of seawater).
Assume a bulk density of 120 pcf, gb = 120 – 64 = 56 pcf.
Assume hmax = 10 ft, a consolidation coefficient for sand of cV = 0.5 ft2/s, and a
drawdown time of Tdraw = 0.25 × TTSU = 0.25 × 20 min × 60 s/min = 300 s. The actual
value for TTSU (here assumed to be 20 min) would be obtained from the ASCE 7-16 “Off-
shore Amplitude and Period” maps for the project location.
Therefore,

 2 hmax  s   2 10  70.4 
  min 0.5,    min 0.5,  
   b cV (Tdraw )    56 0.5(300) 
Λ = min[0.5,1.16] = 0.5

Therefore, the soil effective shear strength will be reduced by a factor of 1 – Λ =


1 – 0.5 = 0.5, or a 50% reduction in strength.
This is the maximum reduction that needs to be considered.
If the preceding example is repeated with gravel having a consolidation coeffi-
cient of cV = 50 ft2/s, then Λ = min[0.5, 0.116] = 0.116, for a shear strength reduction of
only 11.6%.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Foundation Design 191

12.2.3  General Site Erosion


General site erosion is the erosion that occurs in the absence of any obstructions such as
buildings or barriers. Erosion occurs both during inflow and drawdown, although the effect
of pore pressure softening tends to enhance the drawdown scour (Figure G12-16). General
site erosion may be evaluated based on standard literature listed in ASCE 7-16 Commentary,
Section C6.12.2.3 with the effect of pore pressure softening incorporated by multiplying the
buoyant specific weight, gb, of the soil by 1–Λ based on the model of Tonkin et al. (2003)
Downloaded from ascelibrary.org by La Trobe University on 10/20/20. Copyright ASCE. For personal use only; all rights reserved.

(Example G12-2).

Figure G12-16. General scour of coastal plane in Rikuzentakata, Japan. Note partial foundation failure of
the building (circled, N 39.0031; E 141.6254).

Scour need not be considered for rock or other non-erodible materials that are capable
of resisting scour during flow velocities of 30 ft/s or higher.
Channelized scour must also be considered, particularly because of the tendency for
drawdown to focus on low-level existing streams or seawalls and other barriers that might
direct the flow. The effects of pore pressure softening need not be considered during chan-
nelized flow. Figures G12-17 and G12-18 show a deep scour channel that developed between
two large surviving reinforced concrete buildings at the shorefront in Onagawa, Japan,
during the Tohoku tsunami. Survivor video footage shows accelerated flow passing between
the two buildings during drawdown, including significant debris from damaged structures.
Video analysis of the flow in this channel indicates that the velocity exceeded 26 ft/s (8 m/s)
(Chock et al. 2013b).

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


192 Tsunami Loads and Effects
Downloaded from ascelibrary.org by La Trobe University on 10/20/20. Copyright ASCE. For personal use only; all rights reserved.

Figure G12-17. Scour hole caused by channelized debris-laden flow between shorefront Marine Pal
buildings in Onagawa, Japan, during drawdown resulting in increased velocity and significant scour of
roadway between these buildings (N 38.4419; E 141.4473).

Figure G12-18. Roadway and resulting scour pit between the Marine Pal buildings in Onagawa, Japan
(N 38.4419; E 141.4473) (left) before and (right) after the Tohoku tsunami.
Source: Google Earth (left) June 24, 2010, (right) April 5, 2011.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Foundation Design 193

12.2.4  Local Scour


Scour around structures or their foundations can result from sustained flow or plunging scour.

12.2.4.1  Sustained Flow Scour


During both inflow and drawdown, sustained flow develops around structures that survive
the flow loading. This sustained flow leads to enhanced scour particularly around the cor-
Downloaded from ascelibrary.org by La Trobe University on 10/20/20. Copyright ASCE. For personal use only; all rights reserved.

ners of buildings as shown in Figures G12-1 and G12-19.


The scour depth can be computed using dynamic numerical or physical modeling or
empirical methods in the recognized literature. In lieu of more detailed analysis, the scour
can also be estimated based on ASCE 7-16, Table 6.12-1 or ASCE 7-16, Figure 6.12-1, which both
present the same data based on field observations. Figure G12-20 shows the individual field
measurements on which this estimate is based.
If the maximum flow Froude number is less than 0.5, the scour determined from
ASCE 7-16, Table 6.12-1 may be reduced by an adjustment factor that varies linearly from 0 at
the horizontal inundation limit to 1.0 at the point on the transect where the Froude number
is 0.5. This reduction is based on field observations that indicate reduced scour depths as
one approaches the inundation limit.
Field observations show that the extent of scour around an obstruction such as a foot-
ing depends on the soil type (Francis 2008, Chock et al. 2013b). For consolidated or cohesive
soils, the scour should be assumed to extend inward and outward from the structure cor-
ners by a distance equal to the scour depth. For nonconsolidated or noncohesive soils, this
distance must be taken as three times the scour depth. Figures G12-10 and G12-11 illustrate
these two conditions.

Figure G12-19. Localized scour around corners of surviving apartment building with tsunami vertical
evacuation refuge on roof level, Minamisanriku, Japan (N 38.67366; E 141.44537).

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


194 Tsunami Loads and Effects
Downloaded from ascelibrary.org by La Trobe University on 10/20/20. Copyright ASCE. For personal use only; all rights reserved.

Figure G12-20. Observed local scour depth and estimated flow depth for different sediment types, with
bounding plausible design envelopes.
Source: Adapted from Tonkin et al. (2013).

Example G12-3. Scour Depth Example.

Consider the flow depth and Froude number profiles determined for the center tran-
sect at the Monterey, California, prototype building location (Figures G12-21 and G12-
22). Note that the EGLA Froude number is used only for the EGL Analysis, while the
Load Case (LC) 2 Froude number is based on umax and 2/3 × hmax. The LC2 Froude num-
ber drops below 0.5 at a distance of 1,350 ft from the shoreline. Because the project
site is within this distance, the full scour depths given by ASCE 7-16, Table 6.12-1 and
ASCE 7-16, Figure 6.12-1 must be considered. Given that the flow depth at the site is
10.8 ft, which is greater than 10 ft, then the scour depth is 12 ft (3.66 m).
Assuming a project location at 1,400 ft from the shoreline, the flow depth would
be 4.5 ft and the LC2 Froude number would be 0.4. Then the scour depth according to
ASCE 7-16, Table 6.12-1 is 1.2 h = 1.2 × 4.5 = 5.4 ft. However, because the Froude number
is less than 0.5, this scour depth can be reduced by a factor of 0.4/0.5 = 0.8. Therefore,
the scour depth at the second project location is 0.8 × 5.4 = 4.32 ft (1.32 m).

12.2.4.2  Plunging Scour


Plunging scour occurs when the tsunami flow overtops a structure such as a berm, seawall,
or low-rise building. The rapid acceleration of flow on the downslope side of the obstruction
leads to enhanced scour of the soil behind the object (Figures G12-2 and G12-3).
The horizontal extent and depth of plunging scour is to be determined by dynamic
numerical or physical modeling or by empirical methods, or using the procedure described
in the standard. Enhanced scour caused by pore pressure softening need not be considered
when computing plunging scour. This analytical approach by Tonkin et al. (2013) is illustrated
in ASCE 7-16, Figure 6.12-2 and Example@Seismicisolation
G12-4 for a seawall and Example G12-5 for a berm.
@Seismicisolation

Tsunami Loads and Effects


Foundation Design 195
Downloaded from ascelibrary.org by La Trobe University on 10/20/20. Copyright ASCE. For personal use only; all rights reserved.

Figure G12-21. Inundation depth (hi) profile from energy grade line analysis for Monterey, California,
center transect.

Figure G12-22. Froude number profile from energy grade line analysis for Monterey, California, center
transect. @Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


196 Tsunami Loads and Effects

Example G12-4. Plunging Scour over a Seawall.

Consider the cantilever seawall shown in Figure G12-23. The wall height of 10 ft is less
than the maximum tsunami flow depth of 13 ft, resulting in overtopping. This example
demonstrates calculation of the scour depth on the downstream side of the wall, Ds.
Downloaded from ascelibrary.org by La Trobe University on 10/20/20. Copyright ASCE. For personal use only; all rights reserved.

Figure G12-23. Plunging scour caused by flow over a cantilever wall.

The plunging scour depth, Ds is determined using ASCE 7-16

qUsin
Ds  c2 v (6.12-1)
g
where
c2V = Dimensionless scour coefficient, permitted to be taken equal to 2.8;
ψ = Angle between the jet at the scour hole and the horizontal, taken as the
lesser of 75 degrees and the side slope of the overtopped structure on the
scoured side;
g = Acceleration owing to gravity;

U  2 g (h  d d )  2  32.2(13  2)  31.08 ft/s; and


q = Discharge per unit width over the overtopped structure, calculated as

2
q = Cdis 2 g H o3/ 2
3

Ho 3
where Cdis  0.611  0.08  0.611  0.08   0.635
HB 10

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Foundation Design 197

2
so q = 0.635× 2 × 32.2 × 33/ 2 = 17.65ft 2 /s
3

17.65× 31.08× sin 75


Therefore, Ds = 2.8 = 11.4 ft
32.2

This is a significant scour pit, with depth greater than the wall height. Design of
the wall foundation must consider the presence of this scour pit. Similar scour may
Downloaded from ascelibrary.org by La Trobe University on 10/20/20. Copyright ASCE. For personal use only; all rights reserved.

occur on the opposite side of the wall during drawdown if there is no alternate route
for the water to return to the ocean.
In this example, the existing soil level on the inland side of the wall is assumed
to be lower than on the ocean side by a height dd = 2 ft. If the existing soil level on the
inland side is higher than on the ocean side, then dd will be input as a negative value in
the calculation of velocity U. The resulting scour will reduce. For example, assuming
dd = –2 ft gives Ds = 10.5 ft because of the lower velocity of the plunging flow.

Example G12-5. Plunging Scour over a Berm.

Consider the concrete-lined earthen berm shown in Figure G12-24. The berm height
of 10 ft is less than the maximum tsunami flow depth of 13 ft, resulting in overtopping.
This example demonstrates calculation of the scour depth on the downstream side of
the berm, Ds.

Figure G12-24. Plunging scour owing to flow over a berm.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


198 Tsunami Loads and Effects

The plunging scour depth, Ds, is determined using the ASCE 7-16 equation

qUsin
Ds  c2 v (6.12-1)
g

where
c2v = Dimensionless scour coefficient, permitted to be taken equal to 2.8;
Downloaded from ascelibrary.org by La Trobe University on 10/20/20. Copyright ASCE. For personal use only; all rights reserved.

ψ = Angle between the jet at the scour hole and the horizontal, taken as the
lesser of 75 degrees and the side slope of the overtopped structure on the
scoured side (35 degrees);
g = Acceleration owing to gravity; and
q = Discharge per unit width over the overtopped structure, calculated as

2
q = Cdis 2 g H o3/ 2
3

Ho 3
where Cdis  0.611  0.08  0.611  0.08   0.635
HB 10

2
Therefore, q  0.635  2  32.2  33/ 2  17.65 ft 2 /s
3

and U  2 g (h  d d )  2  32.2(13  2)  31.08 ft/s

17.65×31.08×sin 35
Therefore, Ds = 2.8 = 8.75 ft
32.2

This is a significant scour pit, with depth just less than the berm height. Design of
the berm must consider the presence of this scour pit. Similar scour may occur on the
opposite side of the berm during drawdown if there is no alternate route for the water
to return to the ocean.
In this example, the existing soil level on the inland side of the berm is assumed
to be lower than on the ocean side by a height dd = 2 ft. If the existing soil level on the
inland side is higher than on the ocean side, then dd will be input as a negative value in
the calculation of velocity U. The resulting scour will reduce. For example, assuming
dd = –2 ft gives Ds = 8.1 ft because of the lower velocity of the plunging flow.

12.2.5  Horizontal Soil Loads


Unequal scour on opposite sides of a footing or structure will result in unbalanced lateral
earth pressures on opposite sides of the footing or structure (Figure G12-8). This unbalanced
horizontal load must be included in the design of the foundation element.
@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Foundation Design 199

12.2.6  Displacements
Foundation displacements are to be estimated using the same procedures as other geotech-
nical displacement calculations recognized in the literature for the identified cases of foot-
ings, slopes, walls, and piles. The tsunami loads given in ASCE 7-16, Section 6.12.2 are already
in a form consistent with other geotechnical loads.

12.3  Alternative Foundation Performance-Based


Downloaded from ascelibrary.org by La Trobe University on 10/20/20. Copyright ASCE. For personal use only; all rights reserved.

Design Criteria
It is also permitted to use a performance-based geotechnical design methodology to evalu-
ate the overall performance of the foundation system for potential pore pressure softening
by performing a two- or three-dimensional tsunami-soil-structure interaction numerical
modeling analysis. Although ASCE 41 is not strictly appropriate for application to tsunami
conditions, similar methodologies can be utilized. For TRC IV buildings and other struc-
tures, an independent peer review must be performed of the performance-based foundation
design.
For locations where the tsunami may result from a local seismic event, this perfor-
mance-based design must consider the in situ soil and site surface condition at the onset
of tsunami loads, including those existing at the end of the seismic shaking and including
liquefaction, lateral spread, and fault rupture effects.

12.4  Foundation Countermeasures


To protect foundations from the effects of scour, it is permitted to use various ground
improvement or protection measures. These include specially placed structural fill, pro-
tective slab-on-grade, geotextiles and reinforced earth systems, facing systems, or ground
improvement. Observations after past tsunamis have shown that these measures are not
always effective unless they are specifically designed for the tsunami flow conditions antic-
ipated at the site. Slab-on-grade has been observed to fail and roll up, becoming tumbling
debris (Figure G9-4). Protective concrete slabs on the leading and trailing edges of earth-
filled seawalls in Japan were often undercut by the flow and displaced long distances from
their original locations. The resulting scour of the earth fill material often led to complete
failure of many of these walls (Figures G12-25 and G12-26).

12.4.1  Fill
Fill used for structural support and protection shall be placed in accordance with ASCE 24,
Sections 1.5.4 and 2.4.1, which call for structural fill to be placed in lifts of 12 in. loose thick-
ness, with each layer compacted to at least 95% of its maximum standard Proctor density
(ASCE 2005). The fill must be designed to remain stable during inundation and drawdown
given the loads and effects specified in ASCE 7-16, Section 6.12.2.
@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


200 Tsunami Loads and Effects
Downloaded from ascelibrary.org by La Trobe University on 10/20/20. Copyright ASCE. For personal use only; all rights reserved.

Figure G12-25. Failed seawall at Otsuchi Bay, Japan. Seawall consisted of earthen berm covered with
concrete slab lining on both sides (N 39.3382; E 141.8968).
Source: Courtesy of Gary Chock.

Figure G12-26. Concrete lining stripped from earthen berm in Taro, Japan (N 39.7360; E 141.9750).

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Foundation Design 201

12.4.2  Protective Slab-on-Grade


Exterior slab-on-grade must be assumed to fail owing to uplift unless it is specifically designed
to resist the effects of high-velocity tsunami flow. ASCE 7-16 Commentary, Figure C6.12-3
shows the different loading conditions applied to a slab-on-grade during tsunami flow. All
these effects must be considered in the slab design if it is to be considered protective of the
soil substrates. References are given in the ASCE 7-16 Commentary that give guidance for
this design based on roadway design in the coastal environment.
Downloaded from ascelibrary.org by La Trobe University on 10/20/20. Copyright ASCE. For personal use only; all rights reserved.

12.4.3  Geotextiles and Reinforced Earth Systems


Three reinforced earth systems are permitted by the ASCE 7-16 standard, provided design
and installation are in accordance with manufacturer’s installation requirements and as rec-
ommended in the recognized literature. These systems are geotextile tubes, geogrid earth
and slope reinforcement systems, and geocell earth and slope reinforcement systems. The
same resistance factor of 0.67 is applied to the reinforced earth resistance to the various
loading conditions. Broad use of these techniques in coastal environments has proven their
effectiveness.

12.4.4  Facing Systems


Facing systems are permitted for reinforced earth structures. Suitable facing includes geo-
textile filter layers, concrete facing designed in accordance with protective slab-on-grade
countermeasures, stone armoring or riprap designed to resist the tsunami flow, and veg-
etative facing for general erosion and scour resistance where the flow velocities are less
than 12.5 ft/s (3.8 m/s). Figure G12-27 shows an example of the potential uplift pressures on
concrete slabs protecting a soil berm, and Figure G12-28 shows the protection provided by
grass cover on a mound on the Sendai, Japan, plain that was completely overtopped during
the Tohoku tsunami.

12.4.5  Ground Improvement


Soil cement ground improvement for foundations is an effective method to resist erosion
and scour under high-velocity turbulent flows. It is required that the mass strength after
reinforcement be at least 100 psi average unconfined compressive strength. An added ben-
efit of soil cement ground improvement is the reduction of permeability, thereby reducing
the potential for pore pressure softening or seepage. The techniques of deep soil mixing and
jet grouting can be applied to achieve this soil cement ground improvement. Suggested ref-
erences are given in the ASCE 7-16 Commentary.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


202 Tsunami Loads and Effects
Downloaded from ascelibrary.org by La Trobe University on 10/20/20. Copyright ASCE. For personal use only; all rights reserved.

Figure G12-27. Potential uplift of concrete facing slabs on reinforced earth berm.

Figure G12-28. Grass-covered mound on the Sendai, Japan, plain that was completely overtopped
during the Tohoku tsunami (N 38.1724; E 140.95377).

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


13
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/23/20. Copyright ASCE. For personal use only; all rights reserved.

Structural Countermeasures

ASCE 7-16 allows two countermeasures to reduce the structural effects of tsunamis. The first
approach is to design the building as an open structure, allowing for easier flow through the
lower levels of the building, thereby reducing the hydrodynamic lateral loads on the overall
lateral force–resisting system and the individual structural elements. The second is to design
and construct tsunami barriers to reduce or eliminate all tsunami effects on the building
structure.

13.1  Open Structures


Open structures are defined as those “in which the portion within the inundation depth has
no greater than 20% closure ratio, and in which the closure does not include any Tsunami
Breakaway Walls, and which does not have interior partitions or contents that are prevented
from passing through and exiting the structure as unimpeded waterborne debris.” Open
framing, such as the lower levels of the building shown in Figure G13-1, will allow flow to pass
through the building thereby reducing the hydrodynamic drag on the structural members.
For open structures there is no need to check for buoyancy, so Load Case 1 is not required. In
addition, the potential for excessive debris damming is reduced so that the minimum block-
age factor is taken as Ccx = 0.50 instead of the 0.70 required for all other structures. Because
water can flow freely through the ground floor level of the building, the scour requirements
of ASCE 7-16, Section 6.12.2.4 are not required for open structures.
However, construction that leaves this amount of open space unused is generally
expensive and difficult to justify financially. Therefore, the use of tsunami breakaway walls
to enclose the lower levels and turn them into usable space should be considered. These tsu-
nami breakaway walls are distinct from the breakaway walls described in ASCE 7-16, Chapter
5 on flooding. Traditional flood breakaway walls were to be designed to fail when the lateral
pressure reached 10 psf. This would imply that during high-wind events the walls could fail
because of wind pressure even if there was no inundation at the building site. This meant
that these spaces could only be used for limited applications because they were not intended
to survive high-wind or flooding events.

203
@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


204 Tsunami Loads and Effects
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/23/20. Copyright ASCE. For personal use only; all rights reserved.

Figure G13-1. Tsunami vertical evacuation refuge, an example of an open structure, constructed after
the 2004 Sumatra Tsunami in Meuraxa, Banda Aceh, Indonesia.
Source: http://freeimage96.blogspot.com/2013/02/tsunami-escape-building.html

Tsunami breakaway walls are to be designed for the full wind loads required by ASCE 7-16
for the project site. As demonstrated in Example G13-1, the lateral load at the base of an exte-
rior wall owing to hydrostatic pressure from water inundation will exceed the typical lateral
load from wind pressure when the water level is still relatively low. Breakaway walls must be
designed to fail at a predictable load just exceeding that required to resist the design wind
and seismic loads for the given location. However, the addition of tsunami breakaway walls
will mean that the structure no longer qualifies as an open structure according to ASCE 7-16,
so the exemptions allowed for open structures will no longer apply.
Guidance on designing breakaway walls using cold-formed steel studs is given by Klein-
man et al. (2007), and guidance for designing masonry breakaway walls is given by Matsu-
moto and Robertson (2008).

13.2  Tsunami Barriers


Tsunami barriers may be used to protect buildings or other structures from the effects of the
tsunami inundation. The barrier can be designed to prevent any overtopping, in which case
the protected structure need not be designed for tsunami loads and effects. Alternatively,
the protective barrier can be designed to be overtopped during the design event, thereby
reducing the tsunami loading conditions on the protected building or structure. Such barri-
ers will provide complete protection for smaller tsunami events, but must still be designed

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Structural Countermeasures 205

Example G13-1. Tsunami Breakaway Wall Design.

The figure shows a typical ground-floor exterior breakaway wall assumed to be 9 ft


(2.74 m) high. The wall studs span vertically to resist lateral load. It is assumed that
the connection at the base of the studs is designed to break away during the tsunami
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/23/20. Copyright ASCE. For personal use only; all rights reserved.

flow, allowing the wall to swing inward while still attached to the beam above. This will
reduce the likelihood of this wall becoming floating debris.

A typical coastal wind pressure on components and cladding, such as a windward


wall, would be on the order of 38 psf. The shear per unit width at the base of the wall
owing to wind is therefore

1
Vw   38  9  171 lb/ft
2

Assuming an overstrength factor of 2.0 for the breakaway connection, the wall
will break away when the tsunami shear at the base of the wall is VT = 2 × 171 = 342 lb.
Now solve for the water level that will produce this shear force at the base of the
breakaway wall owing to hydrostatic pressure distribution on a unit width of the wall.
Taking moments about the top of the wall

1  1 2  h   1  1 2  h 
VT     s h    9      1.1 64  h    9     342 lb
9  2   3   9  2   3 

Solving gives h = 3.33 ft.


The wall will therefore fail when the water depth is about one-third of the story
height.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


206 Tsunami Loads and Effects

for the effects of overtopping anticipated during the MCT event. This includes the effects of
plunging scour on the back side of the barrier and the effect of trapped water inside the pro-
tected area during drawdown if there are no openings in the barrier. The protected structure
must also be designed for the residual inundation resulting from the MCT event.
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/23/20. Copyright ASCE. For personal use only; all rights reserved.

13.2.1  Information on Existing Buildings and Other


Structures to be Protected
It is permitted to utilize the procedures of ASCE 41, Chapters 9 through 11 to evaluate the
as-built building configuration, building components, site, and foundation. Baiguera et al.
(2020) present an ASCE 41 consistent tsunami non-linear static analysis (push-over) pro-
cedure that complies with ASCE 7-16, Chapter 6 hydrodynamic loading requirements. This
procedure can be applied for evaluation of existing structures and for the design of new struc-
tures according to the alternative performance-based criteria in ASCE 7-16, Section 6.8.3.5.2.

13.2.2  Site Layout


If the tsunami barrier is located close to the protected structure, splash-up can impact the
upper levels of the structure. It is therefore required that the protective barrier be set back
from the structure for perimeter protection. Alignment changes in the barrier shall be grad-
ual so as to reduce the potential for flow acceleration and scour.
If the barrier is not intended to provide protection around the entire building site, then
the extent of the barrier must be sufficient to protect the structure from flow orientations of
+22.5 degrees on either side of the shoreline normal flow direction, as specified in ASCE 7-16,
Sections 6.8.6.1 and 6.8.6.2.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


14
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/23/20. Copyright ASCE. For personal use only; all rights reserved.

Tsunami Vertical Evacuation


Refuge Structures
In locations where evacuation from the tsunami inundation zone to high ground is not possi-
ble, either because of the short warning time or the large population to be evacuated, tsunami
vertical evacuation refuge structures can be used to augment the evacuation options. Fraser
et al. (2012) described the effectiveness of vertical evacuation during the Tohoku tsunami,
particularly because of the short warning time (20 to 30 min after the earthquake shaking)
available to those in the inundation zone. Although many lives were saved by the use of des-
ignated tsunami vertical evacuation refuge structures, tens of thousands of lives were saved
when people sought refuge in buildings that were not designated for vertical evacuation, but
proved to be tall and strong enough to survive without collapse. There were unfortunately
also a number of instances in which designated vertical evacuation refuge structures were
not tall enough to protect the occupants, resulting in an unknown number of deaths. One
particular example was the Emergency Operations Center in Minamisanriku, Japan, (Fig-
ure G14-1). The roof of this structural steel–framed building was the designated refuge for
the 30 emergency responders, including the mayor, who came to the building immediately
after the earthquake. Unfortunately, the tsunami overtopped the roof by as much as 6 ft,
washing away most of those who sought refuge. The mayor and eight others were the only
survivors because they clung to the telecommunication towers and staircase guardrail at
the height of the inundation (Chock et al. 2013b). The skeleton of the building has now been
preserved as a memorial to those who died while serving their community.
Also in Minamisanriku, the Matsubara apartment building was built in 2007 as a desig-
nated vertical evacuation refuge (Figure G14-2). The entire roof of this building served as the
evacuation refuge with capacity for up to 1,000 people. Access to the roof was provided by a
staircase and elevator, and the refuge area was surrounded by a sturdy 6 ft high guard fence.
According to Fraser et al. (2012), 44 people survived the tsunami because they sought refuge on
the roof of this building, even though the water level exceeded the roof elevation by 2 ft (0.7 m).
In Rikuzentakata, Iwate Prefecture, Japan, the second floor of the City Hall Community
Center and Gym was designated as a tsunami refuge (Figure G14-3). Of those who sought

207
@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


208 Tsunami Loads and Effects
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/23/20. Copyright ASCE. For personal use only; all rights reserved.

Figure G14-1. Emergency Operations Center, Minamisanriku, Miyagi Prefecture, Japan (N 38.6778;


E 141.4464), a 3-story structural steel–framed building with roof designated as the evacuation refuge.
The city mayor, Jin Sato, and 29 emergency responders were in the building when the tsunami struck;
24 managed to reach the roof refuge in time. Unfortunately, the refuge elevation was not adequate, and
the guardrail around the refuge failed; only the mayor and eight others survived. The building frame is
now a memorial to the 21 emergency responders who died serving their community.

refuge at this location, nine were members of a swim team at Takata High School who had
evacuated inland from the coast. Only two members of the team survived when the tsunami
inundated the second-floor refuge area (Chock et al. 2013b).
A major lesson from the Tohoku tsunami was therefore the need to ensure that desig-
nated tsunami refuge structures are not only structurally adequate but are tall enough to
ensure safety of the evacuees during the maximum considered tsunami event.
In the United States, FEMA published P-646 Guidelines for Design of Structures for
Vertical Evacuation from Tsunamis, third edition (FEMA 2019). This document is a commu-
nity planning aid that provides guidelines for the siting, funding, and operation of vertical
evacuation refuges. It references ASCE 7-16, Chapter 6 for all tsunami loading and design

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Tsunami Vertical Evacuation Refuge Structures 209
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/23/20. Copyright ASCE. For personal use only; all rights reserved.

Figure G14-2. Matsubara apartment building designated as a tsunami vertical evacuation refuge in


Minamisanriku, Miyagi Prefecture, Japan (N 38.6737; E 141.4454): (a) sign designating roof of building
as “Tsunami Evacuation Area”; (b) large roof area with study 6 ft guard fence; (c) deep pile foundations
supporting structure in spite of significant scour; (d) external elevator and staircase for access to roof
level; (e) foundation design resistant to scour.

Figure G14-3. Rikuzentakata City Hall Community Center and Gym showing damage to second-floor
refuge area (N 39.0139; E 141.6339).

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


210 Tsunami Loads and Effects

requirements. Earlier editions of FEMA P-646 served as valuable prestandards that aided in
the development of the ASCE 7-16, Chapter 6 “Tsunami Loads and Effects” provisions.
This chapter describes the design requirements for a building, mound, or other struc-
ture that will be designated as a vertical evacuation refuge for use during tsunamis. These
provisions are more stringent than those for most other structures in the tsunami design
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/23/20. Copyright ASCE. For personal use only; all rights reserved.

zone because of the potential for loss of life if the designated refuge does not provide safe
haven. Chock et al. (2016) performed a reliability analysis of prototypical buildings designed
to the ASCE 7-16, Chapter 6 requirements. They show that buildings designed as vertical
evacuation refuge structures have less than a 1% probability of failure given that the 2,500-
year maximum considered tsunami has occurred. This compares with a 3% probability of
failure for other Tsunami Risk Category IV buildings and a 7.5% probability of failure for Tsu-
nami Risk Category II buildings given that the MCT has occurred.
To ensure that the refuge elevation is high enough to provide safe haven, it is required
that a site-specific probabilistic tsunami hazard analysis be performed for every tsunami ver-
tical evacuation refuge. This analysis must be performed following the procedure outlined in
ASCE 7-16, Section 6.7 and Chapter 7 of this guide. The water elevation predicted by this PTHA
must then be increased by 30% to accommodate uncertainty in the accuracy of the inundation
modeling. This increase is based on variability in runup model predictions when hindcasting
past tsunami events. For example, Figure G14-4 shows an updated hindcast model based on
Yamazaki et al. (2011) of the Samoa tsunami inundation of Pago Pago harbor on American Samoa
during the September 29, 2009, Samoa tsunami. A number of field-measured runup and flow
elevations exceed the modeled runup elevation by as much as 30%.
To further reduce the probability of overtopping as observed during the Tohoku tsunami,
the refuge areas are to be located one story and at least 10 ft above this amplified water elevation.

14.1  Minimum Inundation Elevation and Depth


To reduce the possibility of overtopping of a designated vertical refuge location during the
maximum considered tsunami, ASCE 7-16 has specific requirements for minimum elevation
of the refuge area. As for all TRC IV structures, site-specific probabilistic tsunami hazard
analysis conforming to the requirements of ASCE 7-16, Section 6.7 is required to determine
the flow depth and velocity at the refuge site. Because of uncertainty in the model predic-
tions of velocity, the PTHA velocity must be compared to that determined using the energy
grade line analysis (ASCE 7-16, Section 6.6) and may not be less than 90% of the EGLA veloc-
ity using urban terrain roughness and not less than 75% of the EGLA velocity for other ter-
rain roughness (see Chapter 6 in this guide).
The flow elevation from the PTHA need not be limited by the flow elevation deter-
mined from EGLA. However, to accommodate potential variability in the actual flow eleva-
tion, a factor of 1.3 must be applied to the flow elevation. It is important to note that this 30%
increase is applied to the elevation above the sea level datum being used, and not the flow
depth at the site. Figure G14-5 shows the difference between these two approaches. The

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Tsunami Vertical Evacuation Refuge Structures 211
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/23/20. Copyright ASCE. For personal use only; all rights reserved.

Figure G14-4. Updated comparison between numerical modeling (red line) and field measurement of
runup (white dots) and flow elevations (blue dots) at Pago Pago Harbor, American Samoa.
Source: Updated version of Yamazaki et al. (2011). Courtesy of Kwok Fai Cheung.

blue dashed line represents the inundation elevation determined from a site-specific PTHA,
drawn along a topographic transect (black line) running perpendicular to the shoreline and
passing through the project site. The flow depth at any location is the difference between
this inundation elevation and the ground elevation. If this flow depth (hmax) were increased
by 30%, it would result in the red dashed line. At the inundation limit, where the runup is
zero, there would be no increase in flow depth and hence no accommodation of the model-
ing uncertainty noted in Figure G14-4. Therefore, it is the flow elevation (z + hmax) that must
be increased by 30% to produce the solid blue line. The resulting flow depth at the project
site is then

hmax = 1.3(z + hmax,ss) – z = 1.3(11.8 + 10.8) – 11.8 = 17.6 ft

This flow depth, along with the flow velocity, is used to determine all hydrostatic, hydro-
dynamic, and impact loads on the structure and to determine the extent of general erosion
and scour around the foundations.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


212 Tsunami Loads and Effects

It should be recognized that evacuation to high ground allows for the possibility of
escape to even higher ground if the tsunami inundation is greater than anticipated. How-
ever, for buildings, mounds, and other artificial high ground used for vertical evacuation,
there may be no higher elevation to which the occupants can escape. To provide an added
measure of safety against potential overtopping of the refuge area, ASCE 7-16 requires that
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/23/20. Copyright ASCE. For personal use only; all rights reserved.

the elevation of the refuge area be one floor, and at least 10 ft, above this flow depth. For
the example shown in Figure G14-5, this would mean the refuge area must be at least 27.6 ft
above the ground elevation, or 39.4 ft above the sea level datum used in the site-specific
PTHA. Example G14-1 demonstrates the determination of the appropriate minimum eleva-
tion for a vertical evacuation refuge.

14.2  Refuge Live Load


Because a tsunami refuge area is anticipated to host a large number of people at the time of
a tsunami warning, it is appropriate to design for assembly live load of 100 psf as required by
ASCE 7-16, Section 6.14.2. This live load is not reducible and must be applied to all accessible
areas within the designated evacuation floor area.

Figure G14-5. Example showing the determination of flow depth resulting from a 30% increase in flow
elevation for use in design of a tsunami vertical evacuation refuge.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Tsunami Vertical Evacuation Refuge Structures 213

Example G14-1. Determination of Vertical Evacuation Refuge Elevation.

Figure G14-6 shows a schematic to illustrate determination of the height requirements


for a tsunami vertical evacuation refuge. If the tsunami is triggered by a near-source
subduction earthquake, then the ground elevation may drop because of co-seismic
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/23/20. Copyright ASCE. For personal use only; all rights reserved.

subsidence. This will result in the sea level changing from the pre-earthquake MHW
level to the dashed black line shown in Figure G14-6. This is the new reference sea
level that must be used in the site-specific PTHA, unless the co-seismic subsidence is
included in the source model of the site-specific PTHA. The grade plane of the struc-
ture (average elevation of finished grade adjoining the structure) is shown as the brown
dashed line. Based on the site-specific PTHA, the inundation elevation is shown by
the blue dashed line. This represents the maximum flow depth from the site-specific
time history at the site, hmax,ss, plus the ground elevation, z. This inundation elevation
is multiplied by 1.3 to get the design inundation shown by the solid blue line. An addi-
tional 10 ft is added to this elevation resulting in the dotted black line. Although the
fourth level is above the design inundation elevation, it may not be used for vertical
evacuation because it does not exceed the minimum 10 ft additional height. For this
example, Floors 5, 6, and the roof can be used as designated vertical evacuation refuge
areas, provided they are accessible at short notice. It is important to include both the
co-seismic subsidence and sea level rise when computing the required elevation of
refuge levels above current MHW datum.

Figure G14-6. Schematic demonstrating the height requirements for a tsunami vertical


evacuation refuge.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


214 Tsunami Loads and Effects

14.3  Laydown Impacts


Survivor videos from past tsunamis show numerous telephone and electrical poles toppled
by the tsunami flow. If these poles are close enough to impact the tsunami vertical evacua-
tion refuge when they fall, then the resulting impact on the structure must be considered in
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/23/20. Copyright ASCE. For personal use only; all rights reserved.

the design of exterior structural members. Pole failure is unlikely to occur when the inun-
dation depth is less than 6 ft, so ASCE 7-16 only requires consideration of laydown impacts if
the maximum inundation depth exceeds 6 ft at the refuge site.

14.4  Information on Construction Documents


For all tsunami vertical evacuation refuge structures, ASCE 7-16, Section 6.14.4 requires that
certain tsunami design criteria be included on the construction documents. The following
list gives a sample of the types of information that should be included:

• Refuge latitude, longitude, sea level datum used in the inundation analysis, and grade
plane elevation above this sea level datum;

• Refuge areas shown on appropriate floor and roof plans, including exiting routes from
each area;

• Occupancy capacity of each refuge area and total refuge capacity; and

• Tsunami design criteria, including

◦ Tsunami model used to perform the site-specific PTHA and the modeler perform-
ing the work,

◦ Manning’s roughness coefficients used in the tsunami inundation modeling,


◦ Whether or not bore conditions are anticipated at the site,
◦ Maximum flow depth at the refuge site based on the site-specific PTHA,
◦ Maximum flow velocity at the refuge site based on the site-specific PTHA,
◦ Maximum flow velocity at the refuge site based on EGLA, and
◦ Maximum flow velocity used in the design.

14.5  Peer Review


Because of the critical nature of vertical evacuation refuge structures and the poten-
tial for significant loss of life if these structures fail or are overtopped during a tsunami,
ASCE 7-16, Section 6.14.5 requires peer review of all refuge designs. The peer reviewer must

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Tsunami Vertical Evacuation Refuge Structures 215

be independent of the engineer of record, must be appropriately licensed, and should have
experience with application of the tsunami design and other provisions of ASCE 7-16. The
peer reviewer is required to prepare a written report for the authority having jurisdiction
over the refuge location as to the design’s compliance with the requirements of ASCE 7-16.
This review should include all aspects of the structural design, including seismic and tsunami
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/23/20. Copyright ASCE. For personal use only; all rights reserved.

loads, to ensure that the structure is indeed capable of serving as a tsunami vertical evacua-
tion refuge in the event of a future tsunami.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/23/20. Copyright ASCE. For personal use only; all rights reserved.

Tsunami Loads and Effects


@Seismicisolation
@Seismicisolation
15
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/23/20. Copyright ASCE. For personal use only; all rights reserved.

Designated Nonstructural
Components and Systems

15.1  Performance Requirements


Many facilities such as ports and harbors are of necessity located within the tsunami inun-
dation zone. Other facilities such as power stations, fuel storage, and wastewater treatment
plants are often located near the ocean for convenience and economic reasons. If these
facilities are designated as critical or essential facilities by the authority having jurisdiction,
then they must be designed to survive the earthquake and tsunami loads and effects without
loss of function. This means that not only the structural components but also all critical non-
structural components and systems must survive the tsunami without damage.
During past tsunami inundation events, damage to critical nonstructural components
has led to loss of function of critical and essential facilities. The most tragic example is the
loss of backup generation capacity at the Fukushima Daiichi Nuclear Power Station in Japan
because of inundation during the Tohoku tsunami. The loss of electrical power and inunda-
tion of many of the backup power generators and electrical distribution equipment meant
that the pumps required to cool the nuclear reactors could not run. Within 4 days of the
tsunami, this loss of cooling water circulation had led to three nuclear reactor meltdowns,
hydrogen-air chemical explosions, and the release of radioactive material that contaminated
the surrounding land and ocean (Wikipedia 2020a). If the critical backup power genera-
tors and related electrical distribution equipment had been elevated above the flow level
or housed in watertight enclosures (as was done at the Fukushima II plant), this cascading
disaster could have been avoided.
A similar but less catastrophic failure occurred during the inundation of Manhattan,
New York, during Hurricane Sandy in 2012. Although many backup generators were located
above the flood level in upper floors of buildings, because of fire code restrictions the fuel
storage for these generators was often housed in the building basement. When the base-
ments flooded, the loss of fuel limited the effectiveness of the backup generators.

217
@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


218 Tsunami Loads and Effects

The Tohoku tsunami completely inundated the Minami-Gamou wastewater treatment


plant located on the shoreline near the Sendai, Japan, harbor (Figure G15-1, left). Although
many of the buildings survived structurally, virtually all the mechanical and electrical equip-
ment critical to the plant operation was damaged beyond repair (Figure G15-1, right). This
plant was the primary wastewater treatment facility for 700,000 inhabitants of Sendai City
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/23/20. Copyright ASCE. For personal use only; all rights reserved.

and surrounding areas. For one month after the tsunami, all wastewater coming into the
plant had to be discharged directly to the ocean until basic primary treatment could be
reinstated.
ASCE  7-16, Section 6.2 defines “Designated nonstructural components and systems”
as those assigned a component importance factor, Ip, equal to 1.5 according to ASCE 7-16,
Section 13.1.3. This includes all nonstructural components and systems that are required to
perform a critical function during and after the earthquake and tsunami, so as to maintain
function of the essential building or facility.
The commentary to ASCE 7-16, Section 6.15 suggests three approaches that can be used
to protect designated nonstructural components:

1. The component can be located above the maximum considered tsunami inundation ele-
vation. Although no reference is made to the 1.3 factor used to establish the tsunami
refuge elevation in Section 6.14, it would make sense to locate designated nonstruc-
tural components as high as possible, but at least 1.3 times the maximum inundation
elevation.
2. The components can be protected from inundation. One option is to surround the facil-
ity with a tsunami barrier designed in accordance with ASCE 7-16, Section 6.13.2. The
barrier is required to have a height at least equal to 1.3 times the maximum inundation
elevation at the site. As an alternative, the components could be protected in water-
tight enclosures. Figure G15-2 shows the protective measures being implemented at
the Minami-Gamou wastewater treatment plant serving Sendai City, Japan, after it was
completely devastated by the Tohoku tsunami. It is important that hydrostatic buoyancy
and all other loading conditions be considered for the design of these enclosures.
3. The components could also be designed specifically for tsunami effects. This could apply
to pipelines and other components that can withstand submersion in water. These com-
ponents would have to be designed to withstand all the hydrostatic, hydrodynamic, and
impact loading conditions required for submerged structural components.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Designated Nonstructural Components and Systems 219
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/23/20. Copyright ASCE. For personal use only; all rights reserved.

Figure G15-1. The Minami-Gamou wastewater treatment plant (left) two days after the Tohoku tsunami;
and (right) an interior view of one of the equipment buildings at this plant (N 38.25; E 141.006).
Source: (Left) Google Earth image (2011); © Digital Globe (2016).

Figure G15-2. New facilities constructed after the Tohoku tsunami with watertight doors at the
Minami-Gamou wastewater treatment plant in Sendai, Japan (N 38.25; E 141.006).

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/23/20. Copyright ASCE. For personal use only; all rights reserved.

Tsunami Loads and Effects


@Seismicisolation
@Seismicisolation
16
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/23/20. Copyright ASCE. For personal use only; all rights reserved.

Tsunami Risk Categories III


and IV Non-Building Structures
Tsunami design requirements for TRC III and IV non-building structures are similar to those
for critical nonstructural components. The structures must either be protected from inun-
dation or designed for all the same tsunami loads and effects required for buildings.

16.1  Requirements for TRC III


Non-Building Structures
For Tsunami Risk Category III non-building structures, ASCE 7-16 allows for the use of tsu-
nami barriers to protect against inundation. The barrier must be designed to satisfy all the
requirements of ASCE 7-16, Section 6.13 and must have a top-of-wall elevation not less than
1.3 times the maximum inundation elevation, as required for tsunami vertical evacuation
refuges.

16.2  Requirements for TRC IV


Non-Building Structures
For Tsunami Risk Category IV non-building structures, ASCE 7-16 allows three options for
maintaining functionality during a tsunami. The designated nonstructural systems in non-
building structures can be

1. Protected from tsunami inundation effects;


2. Elevated above 1.3 times the maximum inundation elevation; or
3. Designed to withstand the effects of tsunami loads in accordance with ASCE 7-16, Sec-
tion 6.8.

221
@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


222 Tsunami Loads and Effects

The barriers used to protect TRC IV non-building structures must be designed to sat-
isfy all the requirements of ASCE 7-16, Section 6.13 and must have a top-of-wall elevation
not less than 1.3 times the maximum inundation elevation. The commentary to ASCE 7-16,
Section 6.16.2 provides an alternative approach in which a protective barrier is provided to
mitigate the inundation depth at the structure, but not prevent inundation completely. The
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/23/20. Copyright ASCE. For personal use only; all rights reserved.

structure will still need to be designed for the reduced flow depths and velocities, and the
barrier will need to consider overtopping effects in addition to all other requirements of
ASCE 7-16, Section 6.13.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


17
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/23/20. Copyright ASCE. For personal use only; all rights reserved.

Frequently Asked Questions

The following frequently asked questions have been collected from various webinars and
presentations made by the author over the past few years.

1. Why is tsunami design only required for Risk Categories III and IV structures?

ASCE 7-16, Section 6.1.1 only requires tsunami design for Tsunami Risk Category III
(critical and high occupancy) and TRC IV (essential and hazardous) buildings and other
structures because failure of these structures represents a major threat to safety and
well-being of the affected communities and surrounding areas. However, a local juris-
diction is encouraged to require tsunami design for taller Tsunami Risk Category II
buildings and other structures so as to increase the availability of vertical refuge sites,
and to improve community resilience and post-tsunami recovery.

2. Why are Risk Category II structures such as single- and multifamily homes not included in
the tsunami design requirements of ASCE 7-16?

Light-framed and low-rise Tsunami Risk Category II structures would be uneconomi-


cal to design for tsunami loads and effects and do not provide any life safety function
during a tsunami because they could be fully submerged. The appropriate response
for occupants of low-rise or light-framed construction is to evacuate to high ground
or designated vertical evacuation structures according to the local tsunami evacuation
plan. Designing low-rise single- and multifamily homes to survive tsunami loads and
effects would give the occupants a false sense of security that they are safe sheltering in
place during a tsunami warning.

223
@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


224 Tsunami Loads and Effects

3. How can a local jurisdiction enact the ASCE 7-16 tsunami provisions for taller Risk Cate-
gory II structures?

If a local jurisdiction wants to adopt the International Building Code and include
ASCE 7-16 tsunami design for Risk Category II buildings, they can use language similar
to the following:
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/23/20. Copyright ASCE. For personal use only; all rights reserved.

SECTION 1615 TSUNAMI LOADS

1615.1 General. The design and construction of Risk Categories III and IV buildings and
structures, and those Risk Category II buildings meeting the criteria of Section 1615.2,
where located in the Tsunami Design Zones defined in the ASCE 7-16 Tsunami Design
Geodatabase (version 2016-1.0), shall be in accordance with Chapter 6 of ASCE 7-16.

1615.2 Criteria for Risk Category II Buildings to be subject to Tsunami-Resilient Design


and Construction. Risk Category II buildings and structures shall comply with Chap-
ter 6 of ASCE 7-16 when meeting all the following conditions:

• Are of Occupancy Classifications A, B, E, I, M, R-1, R-2,1 or Higher Education Laborato-


ries, and

• Located where the tsunami inundation depth is greater than 3 ft (0.914 m) at any loca-
tion within the intended footprint of the structure, and

• The highest occupiable floor exceeds xx height above grade plane and also exceeds the
tsunami inundation depth determined at the site.2

Notes:

1. A – Assembly; B – Business; E – Educational; I – Institutional; M – Mercantile; R-1 –


Transient accommodation; Hotels, Motels; R-2 – Nontransient multiresidence, Apart-
ment, Timeshare, etc.

2. Because roofs are not commonly accessible, evacuees will only be able to access the
highest floor. Tsunami design is only required if the elevation of this floor exceeds the
tsunami inundation depth at the site as determined by ASCE 7-16 procedures.

The height required for the highest occupiable floor will depend on a number of factors
including the building location within the TDZ and the seismic design requirements
for which the building is designed. Guidance for establishing the appropriate building
height threshold is provided in Chock et al. (2018) and Carden et al. (2016).

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Frequently Asked Questions 225

4. Why are Puerto Rico, Guam, American Samoa, and other US territories that are threat-
ened by tsunamis not included in the ASCE Tsunami Design Geodatabase?
During development of the map data for the ASCE Tsunami Design Geodatabase, there
were only sufficient funds to perform the PTHA and inundation modeling for the five US
western states. However, all the provisions in ASCE 7-16, Chapter 6 can be used for tsu-
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/23/20. Copyright ASCE. For personal use only; all rights reserved.

nami design in the US territories and other countries threatened by tsunamis. However,
because offshore wave amplitudes and runup elevations are not provided, this would
require that site-specific PTHA and inundation modeling be performed according to the
requirements of ASCE 7-16, Section 6.7.

5. Can the ASCE 7-16, Chapter 6 provisions be used outside of the United States?
Yes. The tsunami design provisions in ASCE 7-16, Chapter 6 are universally applicable.
However, because offshore wave amplitudes and runup elevations for the 2,500-year
maximum considered tsunami are not provided in the ASCE Tsunami Design Geodata-
base, this would require that site-specific PTHA and inundation modeling be performed
according to the requirements of ASCE 7-16, Section 6.7.

6. How do the tsunami design zone maps relate to evacuation maps?

TDZ maps do not replace existing evacuation maps. TDZ maps are intended purely for
determining whether or not tsunami design is required for a particular structure and for
establishing the flow properties for which the structure must be designed. Evacuation
maps are typically created by the local emergency management agency and are intended
for evacuation planning and execution during a tsunami warning. They are generally
determined based on presumed deterministic tsunamis and not by strict probabilistic
hazard analysis as required for structural design. They are also highly influenced by
social geography, travel routes, and access control considerations. Although the extents
of the two maps may appear similar, they should never be confused or interchanged.

7. How do the tsunami design zone maps relate to Flood Insurance Rate Maps (FIRMs)?
FIRM products are developed for 100-year return period coastal and riverine flooding.
The ASCE 7-16 tsunami design zone maps are developed for 2,500-year return period
tsunami inundation. There is no relationship between the two map products.

8. Why are tsunamis caused by meteor strikes and island flank failures not considered in
development of the probabilistic tsunami offshore wave amplitudes?
These events can produce extreme tsunami waves, but their frequency of occurrence
is extremely low. Without sufficient data to establish a likely return period for a certain
magnitude tsunami from one of these events, it is not possible to include them in a
probabilistic tsunami hazard analysis.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


226 Tsunami Loads and Effects

9. Why are the tsunami design provisions based on a maximum considered tsunami with a
2,500-year return period?

ASCE 7-16 tsunami design is based on an MCT with a 2,500-year return period so as to be


consistent with ASCE 7-16 seismic requirements which utilize an MCE with a 2,500-year
return period for design of essential and critical facilities. For the shorelines of Alaska,
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/23/20. Copyright ASCE. For personal use only; all rights reserved.

Washington, Oregon, and northern California, the design-level tsunami will most likely
be the result of a design-level earthquake along the adjacent subduction zones. It is
therefore logical to consider the same level event for the MCT and MCE.

10. Can the ASCE 7-16 tsunami provisions be used to design for 100-year or 500-year return
period tsunamis?

The ASCE 7-16 provisions are intended only when designing for a MCT with a 2,500-year
return period. To use the provisions for other return periods, new PTHA and inundation
modeling would be required to determine the runup elevations and other parameters
required to establish flow conditions at the project site. However, this would not satisfy
the reliability targets of ASCE 7-16, and use of criteria based on shorter return peri-
ods would result in much higher failure probabilities that would not comply with the
ASCE 7-16 standard.

11. Many tsunamis are generated by near-source earthquakes. How is the damage caused by
the earthquake incorporated into the tsunami design?

Buildings and other structures exposed to near-source earthquakes must be designed


for seismic effects according to the appropriate performance level in ASCE 7-16. For
Risk Categories III and IV structures there will be relatively minor structural damage
after the seismic event, meaning that the majority of the structural capacity is still avail-
able to resist tsunami loads and effects. Although seismic damage to Risk Category II
buildings may be more substantial, the anticipated tsunami performance level is col-
lapse prevention.

12. What is a tsunami bore, and how does it affect structural design?

A tsunami bore forms at the leading edge of many tsunami waves as they approach
shore. A bore resembles a broken wave, but actually it represents the highly turbulent
leading edge of a rapid change in water level, advancing toward and onto land. Tsunami
bores must be considered whenever identified by ASCE 7-16, Section 6.6.4. The pres-
ence of bores results in increased flow velocities, especially near the shoreline, repre-
sented by an increased α term in ASCE 7-16, Equation (6.6-3) of the energy grade line
analysis (ASCE 7-16, Section 6.6.2). Bores also increase the hydrodynamic loads on wide
structural elements such as walls (ASCE 7-16, Section 6.10.2.3) and induce high pres-
sures when trapped in structural wall-slab recesses (ASCE 7-16, Section 6.10.3.3).

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Frequently Asked Questions 227

13. Can the ASCE 7-16, Chapter 6 provisions be used to design tsunami retrofits for existing
buildings?

Although ASCE 7-16 is intended for use in the design of new buildings and other struc-
tures, the provisions in Chapter 6, such as loading expressions and design approach, are
equally applicable to the evaluation of existing buildings. ASCE 7-16, Section 6.8.3.5.2
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/23/20. Copyright ASCE. For personal use only; all rights reserved.

allows for the use of alternative performance-based criteria including linear and non-
linear static analysis, commonly referred to as push-over analysis. These methods could
also be used for assessment of and retrofit design for existing buildings. ASCE 7-16, Sec-
tion 6.13.2.1 also allows for the use of ASCE 41 for evaluating existing buildings. The third
edition of FEMA P-646 (2019) has a new Chapter 9 on “Considerations for Existing Build-
ings” that might also be helpful when assessing existing buildings as possible tsunami
refuges.

14. Can ASCE 7-16, Chapter 6 be used for tsunami design of coastal bridges?

Coastal bridges are designed to AASHTO design specifications, and do not fall under
the purview of ASCE 7-16. Typical return periods for bridge design are 1,000 years, so
the tsunami design zone maps, which are based on a 2,500-year return period, cannot
be used to determine 1,000-year flow conditions at a coastal bridge site. However, the
site-specific PTHA and inundation modeling provisions of ASCE 7-16 could be applied to
determine 1,000-year tsunami flow conditions at a bridge site, and many of the loading
expressions in ASCE 7-16 would also apply to bridge structures. There is a current effort
to introduce tsunami design provisions into the AASHTO bridge design specifications,
in which case those would supersede any application of ASCE 7-16 for bridge design.

15. Can ASCE 7-16, Chapter 6 be used for the design of industrial facilities like fuel storage
tanks and associated piping?

Yes, ASCE 7-16, Chapter 6 can be used for industrial facilities. The tsunami flow parame-
ters and loading expressions can be used for various components of industrial facilities.
There are currently no provisions specific to design of fuel storage tanks; however, the
ASCE 7-16 Tsunami Loads and Effects Subcommittee is currently considering loading
expressions for aboveground piping that may be included in the 2022 version of ASCE 7.

16. What effect does tsunami inundation have on buried utilities?

Significant sediment transport and scour can occur during tsunami inundation and
drawdown. This may expose buried utilities to direct loading from high-velocity flow.
Depending on the soil permeability, it is also possible for buried utilities to experience
buoyancy and overburden effects. Buried fuel storage tanks have suffered uplift and
buckling during past storm surge and tsunami inundation.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


228 Tsunami Loads and Effects

17. How does the built environment affect the tsunami inundation?

The built environment will generally reduce the tsunami flow velocity by providing
increased bottom friction compared with open ground. Light-framed structures such
as single-family and low-rise multiple-family residences often suffer complete failure
during tsunami inundation. They then become waterborne debris that can result in
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/23/20. Copyright ASCE. For personal use only; all rights reserved.

impact loads and debris damming effects on other structures. Structurally substantial
buildings that are able to resist the tsunami loads may result in channeling of the flow,
leading to increased flow velocities for downstream structures. All these effects are
addressed in the ASCE 7-16, Chapter 6 provisions.

18. If a large building is located seaward of my site, can I reduce the tsunami loads owing to
shielding from that building?

ASCE 7-16 does not allow any reduction in tsunami loads owing to the possibility of
shielding from adjacent structures. This mimics the approach used for wind loads, in
which the deleterious effects of adjacent structures must be considered, but no relief
can be taken for shielding by upstream structures, however robust they might be. The
primary reasons for this approach are that the structural integrity of the upstream
structure cannot be relied on to resist the tsunami loads, and future demolition of
upstream structures could expose the subject building to tsunami loads for which it
was not designed.

ASCE 7-16, Section 6.13 does allow for reduction or elimination of tsunami loads using
various protective structural countermeasures. For example, Section 6.13.2 allows for
tsunami barriers to be used as an external perimeter structural countermeasure so long
as they are designed consistent with the protected structure performance objectives to
jointly achieve the performance criteria.

19. When water flows through a narrow opening, it will accelerate. Does Chapter 6 accommo-
date for this effect as the tsunami flows between buildings?

Yes, ASCE 7-16, Section 6.8.5 deals with flow velocity amplification owing to upstream
structures. ASCE 7-16 Commentary, Section C6.8.5 provides some useful research
results that can assist in establishing the effects of upstream structures that channelize
the flow.

20. Why is debris impact not considered for interior columns and structural walls?

During past tsunami events, the majority of floating and tumbling debris strikes have
been observed to impact the exterior of buildings. If a vehicle, log, or shipping container
is able to breach the exterior of the building and get into the interior, it will generally
be traveling at a much reduced speed, so impact loads would be significantly smaller
than considered for exterior structural elements. However, certain structures, such as

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Frequently Asked Questions 229

parking garages, are filled with potential floating debris, so a change proposal is being
considered for the 2022 edition of ASCE 7 to add debris impact loads to interior struc-
tural elements of certain buildings.

21. Why is debris damming not considered for interior columns and structural walls?
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/23/20. Copyright ASCE. For personal use only; all rights reserved.

Observations after past tsunamis have confirmed that large amounts of debris can
accumulate on the exterior of buildings, but generally very little debris is observed to
dam against interior structural elements such as columns and walls. However, recent
tsunami events have demonstrated that the contents of warehouses can become lodged
against interior structural columns and walls. A change proposal is being considered for
the 2022 edition of ASCE 7 to add debris damming for interior structural elements in
warehouses and garages for large vehicles (buses and trucks).

22. Do I need to consider scour around the foundations if the area around the building is cov-
ered by a slab-on-grade?

A typical slab-on-grade will not provide protection when subjected to high-velocity


tsunami flow. During past tsunamis, slabs-on-grade were often uplifted and became
tumbling debris. However, ASCE 7-16, Section 6.12.4 provides various foundation counter-
measures that can be taken to prevent scour. These include a protective slab-on-grade
designed for high-velocity flow, reinforced earth systems, and ground improvement,
among others. ASCE 7-16 Commentary, Section C6.12.4.2 provides additional guidance
on protective slab-on-grade design.

23. What is the difference between NAVD88 and MHW? Why are both provided? Why is only
MHW provided for Hawai‘i and Alaska?

The North American Vertical Datum established in 1988 (NAVD88) is a vertical ref-
erence datum used primarily on land. It was only established for the 48 contiguous
states, so is not available for Hawai‘i and Alaska. The MHW elevation is almost univer-
sally used as the datum for tsunami inundation modeling. The two data differ by vary-
ing amounts along the US western seaboard, often differing by many feet. A change
proposal is currently being considered to utilize MHW as the default datum for tsu-
nami design rather than NAVD88. However, because NAVD88 is commonly used for
building construction it will still be included in the ASCE Tsunami Design Geodatabase
where available.

24. What extent of offshore wave amplitude values should be considered when performing
site-specific PTHA?

Currently the ASCE 7-16 provisions are silent on the extent of offshore wave amplitude
values that must be considered when performing site-specific PTHA. However, a change
proposal is being considered for the 2022 edition of ASCE 7 to provide an extent of

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


230 Tsunami Loads and Effects

offshore wave amplitudes that must be considered so as to ensure hazard consistency


with the 2,500-year probabilistic tsunami.

25. Why are the offshore wave amplitude points located at the 328 ft (100 m) bathymetric line
and not at the shoreline?
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/23/20. Copyright ASCE. For personal use only; all rights reserved.

The offshore wave amplitude points were developed from a probabilistic combination of
thousands of potential tsunami sources around the Pacific basin. This process is greatly
simplified if the waves remain linear throughout propagation from the source to the
integration point. At the 328 ft (100 m) bathymetric depth line all tsunami waves will
still be linear. However, as the waves move into shallower water, they will shoal and
often transform into broken bores before they reach the shoreline. This would require
time-consuming and less reliable nonlinear wave mechanics to estimate the wave ampli-
tudes at or near the shoreline.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

Appendix. Seaside, Oregon:


Full Design Example

A.1  Project Site


The Seaside design example considers a multistory reinforced concrete building in Seaside,
Oregon, at the location shown in Figure GA-1. The center of the building footprint is located
at N 45.99483; W 123.9295, which is 1,025 ft from the shoreline. Figure GA-1 also shows the
three topographic transects along which the energy grade line analysis needs to be applied.
The center transect (C) is drawn perpendicular to the shoreline, represented by the aver-
age coastline for 500 ft either side of the center transect. The clockwise (CW) and coun-
terclockwise (CCW) transects are generated by rotating the center transect through 22.5
degrees in each direction, about the geometric center of the building plan at the grade plane
(ASCE 7-16, Section 6.8.6.1). Each transect is then extended until it reaches the runup points
on the ASCE 7-16 tsunami design zone map. If the end of a transect falls between two of the
runup points, then the runup elevations can be interpolated. The resulting runup eleva-
tions for each transect are given in Table GA-1 along with the approximate inundation limit
distances obtained using the “measure” feature in the ASCE Tsunami Design Geodatabase.
These inundation limit distances will be revised once the runup elevations are plotted on the
respective topographic profiles.
For Washington, Oregon, and California, the ASCE 7-16 TDZ maps provide the runup
elevations in relation to MHW and NAVD88. At the Seaside location the difference between
these two elevation data is 7.1 ft. This difference varies along the Pacific Coast.

239
@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


240 Appendix. Seaside, Oregon: Full Design Example
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

Figure GA-1. Location of the project site in Seaside, Oregon, relative to the inundation line defined by
the ASCE 7-16 tsunami design zone map. The 22.5 degree variation in principal flow direction required
by Section 6.8.6.1 results in clockwise (CW) and counterclockwise (CCW) transects on either side of the
center (C) transect.

Table GA-1. Runup Elevation and Inundation Limits for Three Transects


through the Seaside Project Site

Runup elevation (ft) Inundation limit (ft)

Transect MHW reference NAVD88 reference

Including Including
sea level sea level From TDZ in From
From TDZ rise From TDZ rise Geodatabase transect

Center 48.45 48.55 55.55 55.65 5,640 5,659

Counterclockwise 51.49 51.59 58.59 58.69 6,707 8,168

Clockwise 52.27 52.37 59.37 59.47 7,245 7,168

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Appendix. Seaside, Oregon: Full Design Example 241

A.2  Sea Level Change: Section 6.5.3


ASCE 7-16 Section 6.5.3 requires that any anticipated sea level rise be included in the runup
elevation used in the tsunami design. For this example, we will assume sea level change
based on a 50-year project life cycle. ASCE 7-16 Commentary, Section C6.5.3 provides a link
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

to http://tidesandcurrents.noaa.gov/sltrends for historical sea level trends relative to mean


sea level (MSL).
From the referenced website the following information is obtained:

Hammond, OR 9439011

The mean sea level trend is -1.22 mm/year with a 95% confidence interval of
+/– 1.81 mm/year based on monthly mean sea level data from 1983 to 2014 which is
equivalent to a change of –0.40 ft in 100 years.

The tsunami design should therefore consider the extrapolated prediction of –1.22 +
1.81 = 0.59 mm/year over the 50-year project life cycle. This results in a sea level rise of
29.5 mm or 1.2 in. (0.10 ft). This must be added to the runup elevation for use in the energy
grade line analysis

A.3  Topographic Profiles


The topographic profiles along each of these transects were obtained from the ASCE Tsu-
nami Design Geodatabase. The topographic profiles are shown for the center, counterclock-
wise, and clockwise transects in Figures GA-2, GA-3, and GA-4, respectively. A horizontal
line is plotted on each profile representing the runup elevation (including sea level rise)
for each of these transects relative to the MHW datum from Table GA-1. The point where
this line intersects the profile represents the inundation limit and the starting point for the
energy grade line analysis. The resulting inundation limit should be cross-checked with the
tsunami design zone map inundation line to ensure that they are similar distances from the
shoreline (Table GA-1). If the TDZ inundation is significantly greater than the first intersec-
tion of the runup elevation line with the topographic profile, it may indicate that a region of
high ground is present in the inundation zone. The runup elevation must then be modified to
match this high-ground elevation and the corresponding inundation limit determined where
the modified runup elevation next intersects the topographic profile. The resulting values
for inundation limit are shown in Table GA-1 and are used in the EGLA along each transect.
The project site location is also indicated on each plot. For the center transect, the site
is located 1,025 ft from the shoreline (Figure GA-2). The elevations at the project site vary
slightly for the three transects, which can be attributed to slight differences in the elevation
data points used to generate each transect profile.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


242 Appendix. Seaside, Oregon: Full Design Example
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

Figure GA-2. Topographic profile for center transect.

Figure GA-3. Topographic profile for counterclockwise transect.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Appendix. Seaside, Oregon: Full Design Example 243
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

Figure GA-4. Topographic profile for clockwise transect.

A.4  Tsunami Bore Determination


To determine whether a tsunami bore must be considered at the project site, the conditions
in ASCE 7-16, Section 6.6.4 are evaluated for each transect. Tsunami bores shall be consid-
ered where any of the following conditions exist:

1. Prevailing nearshore bathymetric slope is 1/100 or milder: YES (See Figure GA-5 and
associated discussion).
2. Shallow fringing reefs or other similar step discontinuities: Does not apply.
3. Where historically documented: Does not apply.
4. As described in the recognized literature: Does not apply.
5. As determined by a site-specific inundation analysis: Not required for TRC II buildings.

Therefore, bore loading must be considered in this design.


Figure GA-5 shows the approach used to determine the average nearshore bathymet-
ric slope. A central line is drawn perpendicular to the shoreline. This line is an extension of
the center transect running through the project site. The distance from the shoreline to the
328 ft (100 m) bathymetric line, indicated by the offshore data points in the ASCE Tsunami
Design Geodatabase, is then used to determine the average nearshore bathymetric slope. If
any of the transect lines do not intersect the 328 ft (100 m) bathymetric line, this transect can
be ignored for the purpose of determining whether there is a bore.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


244 Appendix. Seaside, Oregon: Full Design Example
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

Figure GA-5. Determination of the average nearshore slope from the 328 ft bathymetric line to the
shoreline along a line perpendicular to the shoreline and lines rotated 22.5 degrees to either side of the
centerline.

The average nearshore bathymetric slope is then computed using

328
f= in ft
distance

or
100
f= in m
distance

The table in Figure GA-5 shows that the nearshore slope is less than 1/100; therefore,
this project site must consider bores.

A.5  Determination of Inundation Depth and


Flow Velocity Using EGLA
The EGLA is a stepwise procedure starting from the runup elevation at the mapped inunda-
tion limit and working shoreward to get the flow parameters at the site of interest.
A spreadsheet was used to perform this operation along all three transects. A soft-
ware tool, EGL, is available from NumSoft Technologies (2019) to perform the EGL analy-
sis. The input values were the runup, including sea level rise, referenced to MHW datum

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Appendix. Seaside, Oregon: Full Design Example 245

(Table GA-1, third column), the inundation limit distance determined from the topographic
profile (Table GA-1, fifth column), a Manning’s coefficient of 0.030 representing “all other
cases” from ASCE 7-16, Table 6.6-1, and α = 1.3 representing bore conditions at the shoreline
as specified in ASCE 7-16, Section 6.6.4. The resulting inundation depth profiles are shown in
Figure GA-6, Figure GA-7, and Figure GA-8 for the center, counterclockwise, and clockwise
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

transect, respectively. The clockwise transect results in the largest flow depth of 31.4 ft at the
project site, which is the value of hmax that will be used in the subsequent design calculations.
The flow velocity profiles across each transect as determined from the EGLA are shown
in Figures GA-9, GA-10, and GA-11 for the center, counterclockwise, and clockwise tran-
sects, respectively. The minimum flow velocity that may be considered is 10 ft/s, which is
indicated on each of the plots. As with the flow depth, the clockwise transect produces the
largest estimate of flow velocity at 37.92 ft/s, which is the value of umax that will be used in
the design calculations.
All the flow depths and flow velocities determined from the EGLA are listed in Table GA-2.

Figure GA-6. Inundation depth (hi) over topographic transect from energy grade line analysis for the
center transect.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


246 Appendix. Seaside, Oregon: Full Design Example
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

Figure GA-7. Inundation depth (hi) over topographic transect from energy grade line analysis for the
counterclockwise transect.

Figure GA-8. Inundation depth (hi) over topographic transect from energy grade line analysis for the
clockwise transect.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Appendix. Seaside, Oregon: Full Design Example 247
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

Figure GA-9. Flow velocity (ui) profile from energy grade line analysis for the center transect.

Figure GA-10. Flow velocity (ui) profile from energy grade line analysis for the counterclockwise
transect.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


248 Appendix. Seaside, Oregon: Full Design Example
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

Figure GA-11. Flow velocity (ui) profile from energy grade line analysis for the clockwise transect.

Table GA-2. Results of Energy Grade Line Analysis for


Three Transects through Seaside Project Site.

Transect Maximum flow depth, hmax (ft) Maximum flow velocity, umax (ft/s)

Center 26.65 34.34

Counterclockwise 30.44 37.85

Clockwise 31.4 37.92


Note: Bold values are the controlling hmax and umax that must be used in the building design.

A.6  Prototype Concrete Buildings


A.6.1  Six-Story Office Building
As described in Chapter 2, the 6-story office building consists of a special moment-resisting
frame on the perimeter and selected interior frames, and interior gravity columns support-
ing posttensioned floor slabs (Figure GA-12). The lateral framing system has been designed
for wind and seismic requirements of ASCE 7-10 for its location (Table G2-1 and Table G2-2).
The resulting seismic base shear is Eh = 2,435 kips, and the building falls into Seismic Design
Category D. This is a Tsunami Risk Category II building with a mean roof height above grade
plane of 74 ft. With a maximum flow depth of 31.4 ft at the Seaside location, this building could

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Appendix. Seaside, Oregon: Full Design Example 249
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

Figure GA-12. Six-story office building using special reinforced concrete moment frame and post-
tensioned flat slab supported on gravity columns.

function as a “refuge of last resort” at the fourth level (38 ft) up to the roof (if accessible).
This office building cannot be designated as a tsunami vertical evacuation refuge because it
has not been designed for TRC IV and the additional requirements of ASCE 7-16, Section 6.14.

A.6.2  Seven-Story Residential Building


Also introduced in Chapter 2, the 7-story residential building consists of a building frame
system with special reinforced concrete shear walls at exit stairs and elevator core, and inte-
rior gravity columns supporting post-tensioned floor slabs (Figure GA-13). The lateral fram-
ing system has been designed for ASCE 7-10 wind and seismic design criteria appropriate
for the Seaside location (Table G2-1 and Table G2-2). The resulting seismic base shear is Eh =
3,273 kips, and the building falls into Seismic Design Category D. This is a Tsunami Risk Cat-
egory II building with a mean roof height above grade plane of 66 ft. With a maximum flow
depth of 31.4 ft, this building could function as a “refuge of last resort” at the fifth level (39 ft)

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


250 Appendix. Seaside, Oregon: Full Design Example
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

Figure GA-13. Seven-story residential building using special reinforced concrete shear walls and post-
tensioned flat slab supported on gravity columns.

up to the roof (if accessible). This residential building cannot be designated as a tsunami
vertical evacuation refuge because it has not been designed for TRC IV and the additional
requirements of ASCE 7-16, Section 6.14.

A.7  Tsunami Loading Summary


Table GA-3 gives a summary of the tsunami loads determined for the building located at the
selected site. The subsequent sections show detailed calculation of each of these tsunami
loads, along with evaluation of the structural system and components for these loads. These
calculations are more detailed than might be necessary for a typical design project because
the intent here is to provide a complete explanation of the various calculations and their
application.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Appendix. Seaside, Oregon: Full Design Example 251

Table GA-3. Summary of Tsunami Loading for Office and Residential Buildings


at the Seaside Location.

Flow parameters Office building Residential building

  Maximum inundation depth, hmax (ft) 31.4 31.4


Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

  Maximum flow velocity, umax (ft/s) 37.92 37.92

Overall building lateral loading (kips)

  Load Case 1 2,295 2,295

  Load Case 2 7,369 7,369

  Load Case 3 1,226 1,226

Component loading (kips)

  Exterior column hydrodynamic drag 1,298a 1,298a

  Interior column hydrodynamic drag 132.4 110.4

  Exterior column debris impact 107.25b 107.25b

  Exterior wall debris impact — 107.25b

Wall and slab loading (psf)

  Hydrodynamic pressure on walls — 3,163

  Stagnation pressure in the mechanical/ — 1,582c


electrical room

  Surge uplift on elevated slabs — 20


a
Including effect of debris damming, Ccx, applied to column tributary width.
b
Limited by log crushing capacity.
c
Stagnation pressure acting outward on structural walls and floor slab enclosing the mechanical/
electrical room corresponding to the maximum velocity and corresponding flow depth.

A.8  Given Conditions


The following conditions assumed to apply for this example are that the

• Building is oriented with the longitudinal axis parallel to the shoreline;

• Building has no basement;

• Foundation system consists of deep piles with pile caps supporting all shear walls and
all exterior columns;

• Ground-floor slab is not isolated from columns, structural walls, and grade beams;

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


252 Appendix. Seaside, Oregon: Full Design Example

• Top of the first-floor windows is 8 ft above grade, with the windowsill at 3 ft;

• Building location is not in the vicinity of a shipping container storage yard or port
facility, and is therefore not subject to debris impact from shipping containers, ships,
or barges; and
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

• Nonstructural exterior cladding spans vertically between floors.

A.9  Tsunami Design for Office Building


A.9.1  Simplified Equivalent Uniform Lateral
Static Force: Section 6.10.1
In lieu of performing detailed hydrostatic and hydrodynamic analysis, Equation (6.10-1) pro-
vides a simplified but conservative estimate of the maximum lateral load on the building

fuw = 2.5ITSUgsh2max = 2.5 × 1.0 × (1.1 × 64.0) × 31.42 = 173 kips/ft

Assuming Ccx = 0.7 controls according to Section 6.8.7, then F = 0.7 × 254 × 173 = 30,854 kips.
This lateral load can be compared with 0.75 W0 Eh = 0.75 × 3 × 2,435 = 5,479 kips < 30,854 kips.
The detailed analysis for LC2 and LC3 must therefore be performed as shown subsequently.

A.9.2  Overall Building Forces


Section 6.8.3.1 defines three load cases, which must be considered in the design.

A.9.2.1  Load Case 1: Maximum Buoyancy and Associated


Hydrodynamic Drag
The exterior inundation depth need not exceed the lesser of

hext ≤ hmax = 31.4 ft


≤ 14 ft  (ground-floor height)
≤ top of first-story windows = 8 ft CONTROLS

Because the ground floor is not isolated from the structural columns, walls, and grade
beams, any pressure increase below the slab will tend to lift the building. It is conservative to
assume that the soil below the slab is saturated and permeable (e.g., sandy or gravelly subsoil)
so that the groundwater pressure will increase below the building generating the following
buoyancy effect (ASCE 7-16, Section 6.9.1):

Fv = gsVw = (1.1 × 64.0)(254 ft × 88 ft × 8 ft)/1,000 = 12,588 kips (6.9-1)

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Appendix. Seaside, Oregon: Full Design Example 253

Apply load combination 0.9D + FTSU + 1.2 HTSU, where HTSU = 0 because scour is assumed
uniform around the building perimeter, and building dead weight D = 16,000 kips, including
foundation.
Therefore, net uplift = –0.9 × 16,000 + 12,588 = –1,812 kips, downward. Overall uplift is
not sufficient to float the building or result in tension in the deep foundations.
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

In combination with any buoyancy, Load Case 1 requires application of the associated
hydrodynamic drag on the entire building, as follows (ASCE 7-16, Section 6.10.2):

1
Fdx   s ITSU Cd Ccx B(hu 2 ) (6.10-2)
2
where
rs = 1.1 × 2.0 = 2.2 sl/ft3,
ITSU = 1.0 (Table 6.8-1: TRC II),
Cd = 1.4575 (Table 6.10-1 based on B/hsx = 254/8 = 31.8),
Ccx = 1.0 because the exterior walls are assumed to be intact for Load Case 1,
B = 254 ft overall width of building, and
h = 8 ft.

Figure  6.8-1 is used to determine the flow velocity corresponding to an inundation


depth of 8 ft as illustrated in Figure GA-14. For h = 8 ft, h/hmax = 8/31.4 = 0.255. Identifying
this point on the inflow side of Figure 6.8-1(a) indicates that this inundation depth occurs
at t/(TTSU) = 0.068. At the same time in Figure 6.8-1(b) the flow velocity ratio is u/umax = 0.7.
Therefore, the flow velocity is u = 0.7 × 37.92 = 26.54 ft/s.
Therefore, substituting into Equation (6.10-2) gives

Fdx 
1
2
  1
 
 s ITSU Cd Ccx B hu 2   2.2 1.0 1.4575 1.0  254 8  26.542 / 1, 000  2, 295 kips
2
This load is applied as a uniformly distributed exterior pressure on the coastal elevation
of the building over a height of 8 ft above grade. The lateral force–resisting system for the
structure at the first-floor level would be evaluated for this load. The nonstructural exterior
wall should not be designed for this load because it is preferred that nonstructural walls fail
so as to relieve lateral load on the structural frame. Only a portion of this load will go to the
second-floor slab, which therefore has to be resisted by the lateral force–resisting system.
The majority of the load will go directly to the grade beam/foundation system. The entire
lateral load , combined with the uplift due to buoyancy calculated earlier, must be resisted by
the deep foundation assuming scour has already occurred owing to a prior wave with 80% of
the design-level amplitude (ASCE 7-16, Section 6.8.8).
If the ground-floor slab-on-grade were isolated from the building columns, walls, and
grade beams, then any uplift pressures developed below the slab would cause localized slab
failure but would not result in buoyancy of the building.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


254 Appendix. Seaside, Oregon: Full Design Example
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

Figure GA-14. Determining u for LC1 with ASCE 7-16, Figure 6.8-1.

A.9.2.2  Load Case 2: Maximum Flow Velocity


According to Figure 6.8-1, LC2 occurs when the inundation depth is 2/3hmax = 2/3 × 31.4 =
20.93 ft. ASCE 7-16, Section 6.10.2 gives the equation

1
Fdx   s ITSU Cd Ccx B(hu 2 ) (6.10-2)
2

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Appendix. Seaside, Oregon: Full Design Example 255

in which all parameters are the same as for LC1 except Cd = 1.252 (Table 6.10-1 based on B/hsx
= 254/20.93 = 12.14). Because the inundation depth of 20.93 ft exceeds the bottom of the sec-
ond-floor beams (14 ft – 24 in./12) = 12 ft, the inundated area of the beams must be included
in the closure coefficient, which is given by
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

hsx = 20.93 ft,


u = umax = 37.92 ft/s,
hcol EQ = 20.93 ft – 24 in. = 18.93 ft (clear height of submerged moment-resisting frame
columns),
Acol EQ= 18.93 ft × 2.33 ft × 40 = 1,764 ft2 (40 MRF earthquake columns, each 28 in. wide),
hcol Grv= 20.93 ft – 8 in. = 20.26 ft (clear height of submerged gravity load columns),
Acol Grv= 20.26 ft × 2 ft × 16 = 648 ft2 (16 gravity load columns, each 2 ft wide),
Awall = 0 ft2 (no structural walls in this MRF structure), and
Abeam = 24 in. × 254 ft × 1= 508 ft2 (24 in. deep beam for full building width).

Ccx =
∑( A
col + Awall ) + 1.5 Abeam
=
∑((1, 764 + 648) + 0) + 1.5×508 = 0.597 < 0.7
Bhsx 254 ft × 20.93 ft

Therefore, Ccx = 0.7 controls according to Section 6.8.7. So

254 20.93  37.922  


1
2
2

1
2

Fdx   s ITSU Cd Ccx B hu   2.2  1.0  1.252  0.7 
1, 000
 7, 369 kips

This load must be applied as a uniformly distributed exterior load of 7,369/20.93 =


352 kips/ft on the coastal or inland elevation of the building over a height of 20.93 ft above
grade as shown in Figure GA-15. The lateral force–resisting system for the structure at the
first- and second-floor levels would be evaluated for this load. During drawdown, the same
pressure needs to be applied to the inland elevation and the lateral force–resisting system
evaluated for this load.

A.9.2.3  Load Case 3: Maximum Inundation Depth


According to Figure 6.8-1, LC3 occurs when the inundation depth is hmax = 31.4 ft and the flow
velocity is 1/3umax = 1/3 × 37.92 = 12.64 ft/s.
ASCE 7-16, Section 6.10.2, gives the equation

1
Fdx   s ITSU Cd Ccx B(hu 2 ) (6.10-2)
2
in which all parameters are the same as for LC1 except Cd = 1.25 (Table 6.10-1 based on B/hsx =
254/31.4 = 8.1). Because the inundation depth of 31.4 ft exceeds the bottom of the third-floor

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


256 Appendix. Seaside, Oregon: Full Design Example
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

Figure GA-15. LC2 overall tsunami loads on the Seaside office building.

beams (14 ft + 12 ft – 24 in. ) = 24 ft, the inundated area of the second- and third-floor beams
must be included in calculation of closure coefficient, which is given by

hsx = 31.4 ft,


u = 12.64 ft/s,
hcol EQ = 31.4 ft – 24 in. – 24 in. = 27.4 ft (clear height of submerged moment-resisting
frame columns),
Acol EQ= 27.4 ft × 2.33 ft × 40 = 2,557 ft2 (40 MRF earthquake columns, each 28 in. wide),
hcol Grv= 31.4 ft – 8 in. – 8 in. = 30.07 ft (clear height of submerged gravity load columns),
Acol Grv= 30.07 ft × 2 ft × 16 = 962 ft2 (16 gravity load columns, each 2 ft wide),
Awall = 0 ft2 (no structural walls in this MRF structure), and
Abeam = 24 in. × 254 ft × 2= 1,016 ft2 (24 in. deep beam for full building width at second
and third floors).

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Appendix. Seaside, Oregon: Full Design Example 257

Ccx 
( Acol  Awall )  1.5 Abeam

( 2, 557  962   0)  1.5 1016  0.632  0.7
Bhsx 2544 ft  31.4 ft

Therefore, Ccx = 0.7 controls according to Section 6.8.7. So


Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

254 31.4  12.642  


1
2
2

1
2

Fdx   s ITSU Cd Ccx B hu   2.2  1.0  1.25  0.7 
1, 000
 1, 226 kips

This load must be applied as a uniformly distributed exterior load of 1,226/31.4 =


39 kips/ft on the coastal and inland elevations of the building over a height of 31.4 ft above
grade as shown in Figure GA-16. Although LC3 does not control design of the lateral force–
resisting system, the intent of LC3 is to ensure evaluation of components up to the maximum
inundation depth for hydrodynamic load and debris impact.

Figure GA-16. LC3 overall tsunami loads on the Seaside office building.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


258 Appendix. Seaside, Oregon: Full Design Example

A.9.3  Evaluation of the Lateral


Force–Resisting System
Because the structure has been designed for Seismic Design Category D, Section 6.8.3.4 per-
mits the use of 0.75W0Eh to evaluate the LFRS, where Eh is the seismic base shear. From the
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

seismic design of this structure, Eh = 2,435 kips. Therefore,

0.75W0Eh = 0.75 × 3 × 2,435 = 5,479 kips

The controlling load case for overall building tsunami lateral load is LC2, with Fdx = 7,369
kips applied over a height of 20.93 ft. A portion of this load will be resisted by the grade
beam/foundation system as shown in Figure GA-15, reducing the overall load by 2,464 kips.
Therefore, VTSU = 7,369 – 2,464 = 4,905 kips. Applying the LFRS assessment gives

0.75Ω0 Eh = 5, 479 kips > 4, 905 kips ∴ OK

So, the seismic lateral force–resisting system has the capacity to resist the overall tsu-
nami loads and does not need to be strengthened. However, the member forces generated in
the elements of the MRF owing to the overall tsunami lateral load must be combined with the
forces generated by the component loads (hydrodynamic drag or debris impact) acting on
those members. To determine the MRF member forces corresponding to the overall tsunami
lateral load, the building must be analyzed for a seismic base shear, Eh, of

VTSU 4, 905
Eh = = = 2,180 kips
0.75Ω0 0.75× 3

This seismic base shear must be distributed up the height of the building following
ASCE 7-16 seismic design provisions. This Load Case 2 equivalent seismic base shear of 2,180 kips
was applied to the same ETABS model used for the original wind and seismic analysis of the
building. Load Case 3 was also analyzed but did not control any of the member designs. This
ETABS analysis resulted in the column forces shown in Figure GA-17 through Figure GA-19 for
floors one through three, respectively. These systemic forces in each element of the LFRS must
be combined with the member forces generated by component loads on that member.
While acting as part of the lateral force–resisting system, these columns are also sub-
jected to component drag or debris impact loads. According to ASCE 7-16, Section 6.8.3.5, the
columns in the inundated floors must be designed and detailed for these combined forces
“that result from the overall tsunami forces on the structural system combined with any
resultant actions caused by the tsunami pressures acting locally on the individual structural
components for that direction of flow.” All members of the LFRS must resist the forces result-
ing from the overall system analysis, in combination with hydrodynamic and impact loads
acting on the member itself.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Appendix. Seaside, Oregon: Full Design Example 259

Floor 1
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

Figure GA-17. Maximum forces in the first-floor columns caused by the LC2 overall tsunami loads.

Floor 2

Figure GA-18. Maximum forces in the second-floor columns caused by the LC2 overall tsunami load.

Floor 3

Figure GA-19. Maximum forces in the third-floor columns caused by the LC2 overall tsunami load.
@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


260 Appendix. Seaside, Oregon: Full Design Example

A.10  Component Loads


A.10.1  Drag Force on Components: Section 6.10.2.2
A.10.1.1  Exterior Columns
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

For Load Case 1, the exterior cladding is assumed to remain intact. Because the cladding
spans vertically between floors for this example building, none of the hydrodynamic lateral
load in LC1 will be applied directly to the ground-floor columns. (If the exterior cladding
were supported by girts which transferred lateral load to the columns, then the columns
would need to be designed for this load.)
For Load Cases 2 and 3, the exterior columns are assumed to have accumulated debris
resulting in an increased tributary width for hydrodynamic load. Section 6.10.2.2 will require
that Cd = 2.0 and the width dimension, b, be taken as the tributary width multiplied by the clo-
sure ratio value, Ccx, given in Section 6.8.7. Previous calculation of Ccx showed that the default
value of 0.7 controls for LC2 and LC3 for this building; therefore, b = 0.70 × 28 ft = 19.6 ft.
The controlling load case will be LC2, when the inundation depth is he = 20.93 ft and
u = umax = 37.92 ft/s.
The hydrodynamic drag is computed using


19.6 20.93  37.922 

1
2
 2
1
Fd   s ITSU Cd b heu   2.2  1.0  2.0 
2 1, 000
 1, 298 kips (6.10-4)

This load is applied to the column as an equivalent uniformly distributed lateral load of
1,298/20.93 = 62.0 kips/ft over the lower 20.93 ft of the column. The column must be designed
for this load combined with gravity loads using the load combinations in Section 6.8.3.3. In
addition, because the exterior columns are part of the LFRS, these component loads must be
combined with the systemic forces and the column designed for the combined loads.

A.10.1.2  Interior Columns


Interior columns are 24 in. (2 ft) square reinforced concrete columns. For Load Case 1, the
interior is not yet inundated, so there are no hydrodynamic loads on the interior columns.
The controlling load case will be LC2, when the inundation depth is he = 20.93 ft and u = umax
= 37.92 ft/s. ASCE 7-16 does not require that any debris damming be considered for interior
columns; therefore, b = column width = 2.0 ft.
The hydrodynamic drag is computed using

Fd 
1
2
 
 s I tsu Cd b heu 2 (6.10-4)

where Cd = 2.0 for square columns (Table 6.10-2).

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Appendix. Seaside, Oregon: Full Design Example 261

Therefore,


2.0 20.93  37.922
1
2
 2
1
Fd   s ITSU Cd b heu   2.2  1.0  2.0 
2 1, 000
 132 kips

This load is applied to the column as an equivalent uniformly distributed lateral load of
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

132/20.93 = 6.3 kips/ft over the lower 20.93 ft of the column. This load must be combined
with gravity loads using the load combinations in Section 6.8.3.3 and the column capacity
verified.

A.10.2  Other Hydrodynamic Loads


No other hydrodynamic load conditions apply to this building because there are no struc-
tural walls and the spandrel beam is integral with the slab so the lateral load on the beam will
transfer directly to the slab diaphragm.

A.10.3  Debris Impact Loads: Section 6.11


The inundation depth at the site exceeds 3 ft; therefore, exterior structural elements below
the flow depth must be designed for debris impact loads according to Section 6.11.

A.10.3.1  Alternative Simplified Debris Impact Static Load:


Section 6.11.1
In lieu of detailed debris impact analysis, the member can be designed for the maximum
static load given by

Fi = 330CoITSU = 330 × 0.65 × 1.0 = 214.5 kips (6.11-1)

Because the building location is not within an impact zone for shipping containers,
ships, and barges, this value can be reduced by 50%, giving the limiting impact load of

Fi = 0.5 × 214.5 = 107.25 kips

This impact load could be used for the column design without checking the individual
impact load expressions shown. The following calculations are performed for completeness,
and in case they produce a lower impact load than the simplified approach.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


262 Appendix. Seaside, Oregon: Full Design Example

A.10.3.2  Detailed Debris Impact Calculation for Office Building


A.10.3.2.1  Wood Logs and Poles: Section 6.11.2
The nominal maximum instantaneous debris impact force is given by
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

Fni = umax kmd (6.11-2)

where
umax = 37.92 ft/s,
k = EA/L for the wood log with a minimum value of 350 k/in. (4.2 × 106 lb/ft), and
md = 1,000/32.2 = 31.1 sl for the minimum 1,000 lb log.

Therefore,

Fni  umax kmd  37.92 4.2 106  31.1 / 1, 000  433 kips

The design instantaneous debris impact force is then given by

Fi = ITSUCoFni = 1.0 × 0.65 × 433 = 281 kips (6.11-3)

The impulse duration is given by

2md umax 2  31.1 37.92


td    0.0054 s (6.11-4)
Fni 433, 060
The column can be designed using a dynamic analysis by applying a rectangular impulse
with magnitude Fi and duration td. As an alternative, an equivalent elastic static analysis can
be performed of the column subjected to Fi multiplied by a dynamic response factor, Rmax,
given in Table 6.11-1. The ratio of impact duration to natural period of the impacted structural
element is obtained using td and the natural period of the column assumed to be fixed against
rotation at both ends. For this case, the natural period is given by

 L2  
Tcol  2  
 22.373  EI
where
L = Unbraced column length = 14 ft – 24 in. = 12 ft for the ground-floor columns,
ρ = Column mass per unit length = 2.333 ft × 2.333 ft × 150 pcf/32.2 ft/s2 = 25.36 sl/ft,
E = Modulus of elasticity of the column concrete = 3,600 ksi = 518.4 × 106 psf, and
I = Moment of inertia of column section = bd3/12 = 2.333 × 2.3333/12 = 2.47 ft4.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Appendix. Seaside, Oregon: Full Design Example 263

Therefore,

 L2  ρ  122  25.36
Tcol = 2π   = 2π   = 0.00569 s
 22.373  EI  22.373  518.4 ×106 × 2.47
   
The ratio of impact duration to column natural period is therefore td/Tcol = 0.0054/0.00569
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

= 0.95.
Table 6.11-1 gives the dynamic response factor Rmax = 1.75; therefore, the equivalent static
load is given by

Fes = FmaxFi = 1.75 × 281 = 492 kips

This exceeds the maximum required impact force of 107.25 kips; therefore, the column
can be evaluated for a lateral point load of 107.25 kips applied at locations that are critical for
flexure and shear.

A.10.3.2.2  Impact by Vehicles: Section 6.11.3


The impact force is given as Fi = ITSU × 30 = 30 kips. This will not control over the log impact
load as has been determined.

A.10.3.2.3  Impact by Submerged Tumbling Boulder and


Concrete Debris: Section 6.11.4
Because hmax = 31.4 ft > 6 ft, an impact force of Fi = ITSU × 8 = 8 kips shall be applied at 2 ft above
grade. This will not control over the log impact load previously determined.
The debris impact load of 107.25 kips is applied to the exterior columns as a static lateral
load at points critical for flexure and shear, in combination with gravity loads on the column.
It is not combined with other tsunami loads and it need not be applied to interior columns.
However, the forces in the column resulting from the debris impact must be combined with
systemic loads if the member is part of the lateral force–resisting system.
In the event that this load exceeds the column capacity, a detailed debris impact dynamic
analysis can be performed by applying a rectangular impulse with magnitude Fi = 107.25 kips
and duration td = 0.0054 s.

A.11  Column Design for Tsunami Loads


A.11.1  Typical Exterior Column Design
A typical exterior column is chosen at Grid Intersection A-6 (Figure GA-12). The column is
part of the lateral force–resisting system for seismic load designed and detailed for Seismic

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


264 Appendix. Seaside, Oregon: Full Design Example

Design Category D. The column has been designed for gravity and seismic loads resulting in
the cross section shown in Figure GA-20 at the ground-floor level and Figure GA-21 for the
remaining floor levels. The column will now be checked for tsunami load combinations.

A.11.1.1  Column Seismic Design


Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

Seismic design of the columns requires additional column ties to ensure ductility of the yield
zones at each end of the column. These yield zones have a length equal to the maximum
column cross-section dimension, in this case hc = 28 in. The critical shear force in this yield
zone occurs at a distance d from the top and bottom of the column, where d = 28 – 1.5 – 0.5
– 1.27/2 = 25.365 in. The critical shear force for the remaining center section of the column
occurs at d + hc from the end of the column, where d + hc = 25.365 + 28 = 53.365 in. The column
ties required for seismic design must be evaluated for the shear force induced by the tsunami
both in the end section (yield zone) and center section of the column (Figure GA-22).

Figure GA-20. Original exterior column cross section at ground-floor level based on Seismic Design
Category D design.

Figure GA-21. Original exterior column cross section at the second- through sixth-floor levels based on
Seismic Design Category D design.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Appendix. Seaside, Oregon: Full Design Example 265
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

Figure GA-22. Typical exterior column elevation showing end and center sections.

A.11.1.2  Column Gravity Loads


The gravity load tributary area for the typical exterior column is 28 ft by 15 ft. The dead load
at the base of the first-floor column is PD = 395 kips and the reduced live load is PL = 56.5 kips.

A.11.1.3  Column Hydrodynamic Loads


Hydrodynamic loads from Load Case 2 will govern over LC1 and LC3 for this column. Analysis
of the column with the applied LC2 hydrodynamic loads results in the bending moment and
shear force diagrams shown in Figure GA-23.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


266 Appendix. Seaside, Oregon: Full Design Example
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

Figure GA-23. Hydrodynamic loading on exterior column of the Seaside office building owing to Load Case 2.

A.11.1.4  Column Debris Impact Loading


For debris impact loading, a static load of 107.25 kips is applied at all locations that will cause
maximum shear force and bending moments in the column. The maximum shear force in the
yield zone of the column occurs when the debris strikes at a distance d below the beam at
each inundated floor. The maximum shear force in the center section of the column occurs
when the debris strikes a distance d outside the yield zone (i.e., at d + hc). The maximum bend-
ing moments occur when the debris strike is applied at the midheight of the clear column
height. Samples of the resulting shear force and bending moment diagrams for debris impact
at a distance d from the end of the column are shown in Figure GA-24 and Figure GA-25.
A sample of the shear force and bending moment diagrams for debris impact at a distance
d + hc from the end of the first floor column is shown in Figure GA-26. Samples of the shear
force and bending moment diagrams for debris impact at the critical height for bending
moments on first- and third-floor levels are shown in Figure GA-27 and Figure GA-28.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Appendix. Seaside, Oregon: Full Design Example 267
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

Figure GA-24. Impact load applied at distance d away from the end of the column on the ground floor.

Figure GA-25. Impact load applied at distance d away from the end of the column on the third floor.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


268 Appendix. Seaside, Oregon: Full Design Example
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

Figure GA-26. Impact load applied at distance d + hc away from the end of the column on the ground
floor.

Figure GA-27. Impact load applied at critical height for bending moments in the ground-floor column.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Appendix. Seaside, Oregon: Full Design Example 269
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

Figure GA-28. Impact load applied at maximum flow depth representing the worst-case bending
moments for the third-floor column.

A.11.1.5  Column Load Combinations


Table GA-4 summarizes the maximum axial load, bending moment, and shear forces for the
inundated levels of the exterior column. These loads are combined using the load combi-
nations provided in Section 6.8.3.3 for both hydrodynamic drag (hydro) and debris impact
(impact) on the column. In addition, because this exterior column is part of the LFRS,
Table GA-4 also lists the maximum axial load, bending moment, and shear forces determined
by the ETABS analysis for the equivalent tsunami base shear (overall) (see Section A.9.3). These
overall systemic forces are then combined with the controlling component forces (hydro or
impact) to obtain the combined forces. It is evident for this exterior column that the hydro-
dynamic drag, including the effects of debris damming, far exceeds the forces induced by the
impact load.
These load combinations are performed for all exterior columns in the building. The
original column designs are then evaluated for these tsunami load combinations and modi-
fied if necessary.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


270 Appendix. Seaside, Oregon: Full Design Example

Table GA-4. Load Combinations for the Seaside Office Building Exterior Column

Moment Axial load Shear at d Shear at d + hc


(k-ft) (kips) (kips) (kips) Load combination

Floor 1
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

1,029 502 307 162 1.2D+FTSU+0.5L (hydro)

1,029 356 307 162 0.9D+FTSU (hydro)

227 502 88 61 1.2D+FTSU+0.5L (impact)

227 356 88 61 0.9D+FTSU (impact)

701 334 85 85 1.2D+FTSU+0.5L (overall)

701 188 85 85 0.9D+FTSU (overall)

1,526 398 392 247 1.2D+FTSU+0.5L (combined)

1,526 252 392 247 0.9D+FTSU (combined)

Floor 2

831 419 220 75 1.2D+FTSU+0.5L (hydro)

831 296 220 75 0.9D+FTSU (hydro)

221 419 88 60 1.2D+FTSU+0.5L (impact)

221 296 88 60 0.9D+FTSU (impact)

651 296 113 113 1.2D+FTSU+0.5L (overall)

651 173 113 113 0.9D+FTSU (overall)

1,007 336 333 188 1.2D+FTSU+0.5L (combined)

1,007 213 333 188 0.9D+FTSU (combined)

Floor 3

140 335 15 15 1.2D+FTSU+0.5L (hydro)

140 237 15 15 0.9D+FTSU (hydro)

221 335 87 60 1.2D+FTSU+0.5L (impact)

221 237 87 60 0.9D+FTSU (impact)

117 331 20 20 1.2D+FTSU+0.5L (overall)

117 233 20 20 0.9D+FTSU (overall)

328 331 107 80 1.2D+FTSU+0.5L (combined)

328 233 107 80 0.9D+FTSU (combined)

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Appendix. Seaside, Oregon: Full Design Example 271

A.11.1.6  Exterior Column Design for Combined Gravity and


Tsunami Loads
The column at Grid Intersection A-6 in Figure GA-12 will now be checked for combined
flexure and axial loads attributed to gravity and tsunami load combinations defined in Sec-
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

tion 6.8.3.3. Figure GA-29 shows the interaction diagram for the exterior column including
the tsunami load combinations. The blue solid line (original column design) represents the
design strength for the original column. The hollow symbols are the tsunami load combina-
tions considering only the hydrodynamic drag force on the individual column component.
Because this column is part of the LFRS, the effects of the component loads must be added to
those of the overall system loads, resulting in the solid symbols shown in Figure GA-29. The
red dashed line (new column design) represents the design strength required to resist the
overall systemic loading combined with the component hydrodynamic drag acting on this
exterior column. Figure GA-30 shows the new column design interaction diagram with the
controlling load combinations for all exterior columns at the ground level. This new column
design is therefore adequate for all exterior columns on the ocean and inland sides of the
building.
Figure GA-31 shows the original and new column interaction diagrams for Column A-6 at
the second-floor level, along with the same loading combinations as described. At the third-
floor level, the original column is adequate to resist the tsunami loads with no modifications.

Figure GA-29. Interaction diagram for typical ground-floor exterior column showing tsunami load
combinations.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


272 Appendix. Seaside, Oregon: Full Design Example
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

Figure GA-30. Interaction diagram for typical ground-floor exterior column showing all combined
tsunami load combinations.

Figure GA-31. Interaction diagram for typical second-floor exterior column showing tsunami load
combinations.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Appendix. Seaside, Oregon: Full Design Example 273

A.11.1.7  New Typical Exterior Column Design


Based on the interaction diagrams shown in Figure GA-29 through Figure GA-31 the original
exterior columns at the ground floor (Figure GA-20) and second floor (Figure GA-21) must be
strengthened to resist bending caused by the hydrodynamic component and overall system
loads. The revised column designs are shown in Figure GA-32 for the ground-floor columns
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

and Figure GA-33 for the second-floor columns. To economize on formwork, the same 28 in.
× 28 in. column was maintained, but the reinforcing increased. The ties in these columns
are designed for seismic ductility requirements and checked for the applied tsunami shear
forces as shown in Section A.11.1.8

Figure GA-32. Exterior column cross section at ground-floor level based on tsunami design
requirements.

Figure GA-33. Exterior column cross section at the second-floor level based on tsunami design
requirements.

A.11.1.8  Exterior Column Shear Design


For critical shears in column at the ground floor, at the critical axial load combination of (0.9D
+ FTSU) according to Equation (6.8-1b), Pu = 251.5 kips. The shear capacities of the 28 in. × 28 in.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


274 Appendix. Seaside, Oregon: Full Design Example

columns with four leg #5 ties at 4 in. on center in the end section and four leg #4 ties at 5 in.
on center in the center section are given by

ϕVn = ϕ(VC + VS)

 
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

Pu  251, 500 
where VC = 2 f C ' 1   bd = 2 4, 000 1   28  25.295 / 1, 000 = 104 kips.
 2, 000 Ag  2, 000  28  28 
 

In the yield zones

Av f y d  4  0.31  60, 000  25.295


VS = = = 470 kips
s 4 1, 000
In the center section

Av f y d  4  0.2   60, 000  25.295


VS = = = 243 kips
s 5 1, 000
Therefore, in the yield zones ϕVn = 0.75 (104 + 470) = 430 kips and in the center section
ϕVn = 0.75 (104 + 243) = 260 kips.
At a distance d from the ends of the column, Vu = 392 kips < φVn = 430 kips; therefore,
the column is adequate for shear in the yield zone.
At a distance d + hc from the column ends, Vu = 247 kips ≤ φVn = 260 kips; therefore, the
column is adequate for shear in the center section.
For critical shears in columns at the second floor, at the critical axial load combination
of (0.9D + FTSU) according to Equation (6.8-1b), Pu = 213.25 kips. The shear capacities of the
28 in. × 28 in. columns with four leg #5 hoops at 4 in. on center in the yield zone and four leg
#4 hoops at 6 in. on center in the center section are given by

ϕVn = ϕ(VC + VS)


 Pu   213, 250 
where VC = 2 f C ' 1   bd = 2 4, 000 1   28  25.295 / 1, 000 = 102 kips.
 2, 000 Ag  2, 000  28  28 
 

In the yield zone

Av f y d  4  0.31  60, 000  25.295


VS = = = 470 kips
s 4 1, 000
In the center section

Av f y d  4  0.2   60, 000  25.295


VS = = = 202 kips
s 6 1, 000

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Appendix. Seaside, Oregon: Full Design Example 275

At a distance d from the ends of the column, Vu = 333 kips < φVn = 0.75(102 + 470) =
429 kips; therefore, the column is adequate for shear in the yield zone.
At a distance d + hc from the ends of the column, Vu = 188 kips < φVn = 0.75(102 + 202) =
228 kips; therefore, the column is adequate for shear in the center section.
For critical shears in exterior columns at the third floor, at the critical axial load combi-
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

nation of (0.9D + FTSU) according to Equation (6.8-1b), Pu = 170 kips. The shear capacities of the
original 28 in. × 28 in. columns with three leg #4 hoops at 4 in. on center in the yield zones
and three leg #3 hoops at 6 in. on center in the center section are given by

ϕVn = ϕ(VC + VS)


 Pu   
where VC = 2 f C ' 1 +
170, 000
 bd = 2 4, 000 1 +  28× 25.436 / 1, 000 = 100 kips.
 2, 000 Ag   2, 000 × 28× 28 

In the yield zones

Av f y d  3  0.2   60, 000  25.436


VS = = = 229 kips
s 4 1, 000
In the center section

Av f y d  3  0.11  60, 000  25.436


VS = = = 84 kips
s 6 1, 000
At a distance d from the ends of the column, Vu = 107 kips < φVn = 0.75(100 + 229) =
247 kips; therefore, the column is adequate for shear in the yield zones.
At a distance d + hc from the ends of the column, Vu = 80 kips < φVn = 0.75(100 + 84) =
138 kips; therefore, the column is adequate for shear in the center section.

A.11.2  Typical Interior Column Design


A typical interior column is chosen at Grid Intersection B-3 in Figure GA-12. The column
is not part of the lateral force–resisting system for seismic loads. It will be subject to the
same displacements as the lateral resisting system and needs to be detailed accordingly for
Seismic Design Category D. It is assumed to have a fixed base at the foundation; therefore,
a plastic hinge may form at the base of the column. The remainder of the column will not
form a plastic hinge, but the slab-column connections need to be detailed appropriately for
the seismic deformation. The 24 in. square column cross-section shown in Figure GA-34 was
selected based on gravity load design and punching shear capacity of the floor slabs.
The critical shear force for the end section of the column occurs at a distance, d, from
the ends of the column, where d = 24 – 1.5 – 0.5 – 0.5 = 21.5 in. The critical shear force for the
center section of the column occurs at d + hc from the end of the column, where d + hc = 21.5
+ 24 = 45.5 in.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


276 Appendix. Seaside, Oregon: Full Design Example
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

Figure GA-34. Interior column cross section at all floor levels based on Seismic Design Category D design.

A.11.2.1  Column Gravity Loads


The gravity load tributary area for the typical interior column is 28 × 29 ft. The dead load at
the base of the first-floor column is PD = 629 kips, and the reduced live load is PL = 114 kips.

A.11.2.2  Column Hydrodynamic Loads


Hydrodynamic loads from Load Case 2 will govern over LC1 and LC3 for this column. Analysis
of the column with the applied LC2 hydrodynamic loads results in the bending moment and
shear force diagrams shown in Figure GA-35.

Figure GA-35. Hydrodynamic loading on interior column of the Seaside office building owing to Load
Case 2.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Appendix. Seaside, Oregon: Full Design Example 277

A.11.2.3  Column Load Combinations


Table GA-5 summarizes the maximum axial load, bending moment, and shear forces for all
inundated levels of the interior columns using the load combinations provided in Section
6.8.3.3 for the hydrodynamic drag (hydro). ASCE 7-16 does not require design for debris
impact on interior columns. The original column design will be evaluated for these load com-
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

binations and modified if necessary.

Table GA-5. Load Combinations for the Seaside Office Building Interior Column.

Moment Axial load Shear at d Shear at d + hc


(k-ft) (kips) (kips) (kips) Load combination

Floor 1

87 812 28 17 1.2D+FTSU+0.5L (hydro)

87 566 28 17 0.9D+FTSU (hydro)

Floor 2

71 677 20 10 1.2D+FTSU+0.5L (hydro)

71 472 20 10 0.9D+FTSU (hydro)

Floor 3

12 541 1 1 1.2D+FTSU+0.5L (hydro)

12 377 1 1 0.9D+FTSU (hydro)

A.11.2.4  Interior Column Evaluation for Combined Gravity and


Tsunami Loads
Figure GA-36 shows the interaction diagram for a typical interior ground level-column with
the tsunami load combinations. The existing interior column is adequate at the first-floor
level, and by inspection the remaining interior columns are also adequate to resist the tsu-
nami bending moments.

A.11.2.5  Interior Column Shear Design


For critical shears in interior columns at the first floor, at the critical axial load combination
of 1.2D + FTSU + 0.5L according to Equation (6.8-1a), Pu = 812 kips. The shear capacities of the

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


278 Appendix. Seaside, Oregon: Full Design Example
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

Figure GA-36. Interaction diagram for typical office building interior ground-floor column showing
tsunami load combinations.

existing 24 × 24 in. column with three leg #4 ties at 4 in. on center in the yield zone and three
leg #3 ties at 5 in. on center in the rest of the column are given by

ϕVn = ϕ(VC + VS)


 Pu   812, 000 
where VC = 2 f C ' 1 +  bd = 2 4, 000 1 +  24 × 21.5 / 1, 000 = 111 kips
 2, 000 Ag   2, 000 × 24 × 24 

In the yield zone


Av f y d  3  0.2   60, 000  21.5
VS = = = 194 kips
s 4 1, 000
In the rest of the column

Av f y d  3  0.11  60, 000  21.5


VS = = = 85 kips
s 5 1, 000
Therefore, in the yield zone, ϕVn = 0.75 (111 + 194) = 229 kips; and in the center sections,
ϕVn = 0.75 (111 + 85) = 147 kips.
At a distance d from the end of the column, Vu = 43 kips < φVn = 229 kips; therefore, the
column is adequate for shear in the yield zone. At a distance d + hc from the end of the col-
umn, Vu = 26 kips < φVn = 147 kips; therefore, the column is adequate for shear at all sections.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Appendix. Seaside, Oregon: Full Design Example 279

A.12  Tsunami Design for Residential Building


A.12.1  Simplified Equivalent Uniform Lateral
Static Force: Section 6.10.1
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

In lieu of performing detailed hydrostatic and hydrodynamic analysis, Equation (6.10-1) pro-
vides a simplified but conservative estimate of the maximum lateral load on the building

fuw  2.5 ITSU  s hmax


2
 2.5 1.0  1.1 64.0   31.42  173.53 kips/ft

Assuming Ccx = 0.7 controls according to Section 6.8.7, then F = 0.7 × 254 × 173.53 =
30,854 kips.
This lateral load can be compared with 0.75 W0 Eh = 0.75 × 2.5 × 3,273 = 6,137 kips <
30,854 kips. The detailed analysis for LC2 and LC3 must therefore be performed as shown in
the following.

A.12.2  Overall Building Forces


Section 6.8.3.1 defines the following three load cases, which must be considered in the design.

A.12.2.1  Load Case 1: Maximum Buoyancy and Associated


Hydrodynamic Drag
The ground floor of the residential building will typically consist of a slab-on-grade. This
slab is assumed to be isolated from the building columns, structural walls, and foundations,
so that any uplift pressures developed below the slab will cause localized slab failure but will
not result in buoyancy of the building. Therefore, overall buoyancy is not a concern. Load
Case 1 also requires application of the associated hydrodynamic drag on the entire building.
This is identical to the corresponding calculation for the office building (Section A.9.2.1),
resulting in an overall hydrodynamic load of Fdx = 2,295 kips. This load would be applied as a
uniformly distributed exterior pressure on the coastal elevation of the building over a height
of 8 ft above grade. The lateral force–resisting system for the structure would be evaluated
for this load. The nonstructural exterior wall should not be designed for this load because it
is preferred that nonstructural walls fail to relieve lateral load on the structural frame. Only a
portion of this load will go to the second-floor slab, which therefore has to be resisted by the
lateral force–resisting system. The majority of the load will go directly to the grade beam/
foundation system. The entire lateral load must be resisted by the deep foundation assuming
scour has already occurred because of a prior wave with 80% of the design level amplitude
(ASCE 7-16, Section 6.8.8).

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


280 Appendix. Seaside, Oregon: Full Design Example

A.12.2.2  Load Case 2: Maximum Flow Velocity


According to Figure 6.8-1, LC2 occurs when the inundation depth is 2/3hmax = 2/3 × 31.4 =
20.93 ft. Section 6.10.2, Equation (6.10-2) gives

 
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

1
Fdx   s ITSU Cd Ccx B hu 2
2
in which most of the parameters are the same as for LC2 for the office building (Section
A.9.2.2).
Because the inundation depth of 20.93 ft exceeds the bottom of the third-floor slab
(12 ft + 9 ft – 8 in./12) = 20.33 ft, the inundated edge area of (1 + 0.6/0.67) = 1.9 floor slabs must
be included in the closure coefficient, which is determined as follows:

hsx = 20.93 ft,


u = umax = 37.92 ft/s,
Acol = 1.67 ft × (20.93 ft – 1.9 × 0.67 ft) × 32 = 1,050 ft2,
Awall = 2 × 28 ft × (20.93 ft – 1.9 × 0.67 ft) + 2 × 10 ft × (20.93 ft – 1.9 × 0.67 ft) = 1,494 ft2,
Abeam = Aslab = 1.9 × 254 ft × 0.67 ft = 323 ft2, and

Ccx 
( Acol  Awall )  1.5  Abeam

 1, 050  1, 494   1.5  323  0.57  0.7
Bhsx 254  20.93

Therefore, Ccx = 0.7 controls according to Section 6.8.7. So

1 1 254(20.93× 37.92 2
)
Fdx = rs ITSU Cd Ccx B (hu 2 ) = × 2.2 ×1.0 ×1.252 × 0.7 × = 7, 369 kips
2 2 1, 000

As for the office building, this load must be applied as a uniformly distributed exterior
load of 7,369/20.93 = 352 kips/ft on the coastal or inland elevation of the building over a
height of 20.93 ft above grade as shown in Figure GA-37. The lateral force–resisting system
for the structure at the first two floor levels would be evaluated for this load. During draw-
down, the same pressure needs to be applied to the inland elevation and the lateral force–
resisting system evaluated for this load.

A.12.2.3  Load Case 3: Maximum Inundation Depth


Determination of the lateral load owing to LC3 is identical to that shown previously for the
office building (Section A.9.2.3). The resulting lateral load of Fdx = 1,226 kips must be applied
as a uniformly distributed exterior load of 1,226/31.4 = 39 kips/ft separately on the coastal
and inland elevations of the building over a height of 31.4 ft above grade.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Appendix. Seaside, Oregon: Full Design Example 281
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

Figure GA-37. LC2 overall tsunami loads on the Seaside residential building.

A.12.2.4  Evaluation of Lateral Force–Resisting System


Because the structure has been designed for Seismic Design Category D, Section 6.8.3.4 per-
mits the use of 0.75W0Eh to evaluate the LFRS, where Eh is the seismic base shear. From the
seismic design of this structure, Eh = 3,273 kips. Therefore,

0.750 Eh  0.75  2.5  3, 273  6,137 kips

The controlling load case for overall building tsunami lateral load is LC2, with Fdx =
7,368 kips applied over a height of 20.93 ft. A portion of this load will be resisted by the grade
beam/foundation system as shown in Figure GA-37, reducing the overall load by 2,112 kips.
Therefore, VTSU = 7,368 – 2,112 = 5,256 kips. Applying the LFRS assessment gives

0.750 Eh  6,137 kips  5, 256 kips  OK

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


282 Appendix. Seaside, Oregon: Full Design Example

Therefore, the special structural shear walls that make up the lateral force–resisting
system are adequate to resist the tsunami loads on the overall building without requiring
strengthening. These walls and the gravity load columns must still be evaluated for compo-
nent loading caused by hydrodynamic drag and debris impact.
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

A.13  Component Loads


A.13.1  Drag Force on Components: Section 6.10.2.2
A.13.1.1  Exterior Columns
For Load Cases 2 and 3, the exterior columns are assumed to have accumulated debris
resulting in an increased tributary width for hydrodynamic load. Section 6.10.2.2 requires
that Cd = 2.0 and the width dimension, b, be taken as the tributary width multiplied by the
closure ratio value, Ccx, given in Section 6.8.7. Therefore, b = 0.70 × 28 ft = 19.6 ft. This is the
same tributary width as for the office building, so again the controlling load case will be LC2,
when the inundation depth is he = 20.93 ft and umax = 37.92 ft/s.
The hydrodynamic drag is computed using Equation (6.10-4) as

Fd 
1
2
  1
 
 s ITSU Cd b heu 2   2.2 1.0  2.0 19.6 20.93  37.922 / 1, 000  1, 298 kips
2
This load is applied to the column as an equivalent uniformly distributed lateral load of
1,298/20.93 = 62 kips/ft over the lower 20.93 ft of the column. The column must be designed
for this load combined with gravity loads according to Section 6.8.3.3.

A.13.1.2  Interior Columns


Interior columns are 20 in. (1.67 ft) square reinforced concrete columns. The controlling load
case will be LC2, when the inundation depth is he = 20.93 ft and umax = 37.92 ft/s. ASCE 7-16
does not require that any debris damming be considered for interior columns; therefore, b =
column width = 1.667 ft. The hydrodynamic drag is computed as

Fd 
1
2

 s ITSU Cd b heu 2  (6.10-4)

where Cd = 2.0 for square columns (Table 6.10-2).

Therefore,

1 1 1.667 (20.93× 37.92 2


)
Fd = rs ITSU Cd b (he u 2 ) = × 2.2 ×1.0 × 2.0 × = 110.4 kips
2 2 1, 000

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Appendix. Seaside, Oregon: Full Design Example 283

This load is applied to the column as an equivalent uniformly distributed lateral load of
110.4/20.93 = 5.27 kips/ft over the lower 20.93 ft of the column. This load must be combined
with gravity loads using the load combinations in Section 6.8.3.3 and the column capacity
verified.
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

A.13.2  Tsunami Loads on Structural Walls, Fw:


Section 6.10.2.3
Because tsunami bores are anticipated at this location, the lateral load on the structural walls
is given by Equation (6.10-5a) or Equation (6.10-5b), depending on the flow depth relative to
the wall width

FW 
1
2

 s ITSU Cd b heu 2  (6.10-5a)

b
FW 
3
4
 
 s ITSU Cd b heu 2 when w ≥ 3
he
(6.10-5b)

where Cd is 2.0 for a wall according to Table 6.10-2.


For the elevator and mechanical room walls, bw = 28 ft. For Load Case 2, which rep-
resents the controlling hydrodynamic conditions

bw 28
= = 1.34  3
he 20.93
Equation (6.10-5a) applies. Therefore,

1
Fd = × 2.2 ×1.0 × 2.0 × 28(20.93× 37.922 ) / 1, 000 = 1, 854 kips
2

This load is applied to the wall as a uniformly distributed pressure of 1,854/(28 × 20.93)
= 3,163 psf over the lower 20.93 ft of the wall.
It is possible that the inundation occurs as a series of bores each with height less than
the flow depth for LC2. In this case, a critical bore height would be

he 1 b 28 ft
= → he = w = = 9.33 ft
bw 3 3 3

because this would require consideration of Equation (6.10-5b). For a flow depth of 9.33 ft,
the flow velocity can be obtained from ASCE 7-16, Figure 6.8-1 (Figure GA-38). Entering Fig-
ure 6.8-1(a) with h/hmax = 9.33 ft/31.4 ft = 0.297 on the inflow side indicates that this inunda-
tion depth occurs at t/(TTSU) = 0.08. At the same time in Figure 6.8-1(b), the flow velocity ratio
is u/umax = 0.78. Therefore, the flow velocity is u = 0.78 × 37.92 = 29.6 ft/s. The bore loading is

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


284 Appendix. Seaside, Oregon: Full Design Example

computed using Equation (6.10-5b) with he = 9.33 ft and u = 29.6 ft/s. Therefore, for the 28 ft
wide elevator wall,

3
Fd = × 2.2 ×1.0 × 2.0 × 28(9.33× 29.62 ) / 1, 000 = 755 kips
4
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

This load is applied to the wall as a uniformly distributed pressure of 755/(28 × 9.33) =
2,890 psf over the lower 9.33 ft of the walls. This will not govern when compared with the
3,163 psf pressure from Equation (6.10-5a). The 28 ft wide elevator and mechanical room
walls must be evaluated for this pressure.
For the stairwell walls, bw = 10 ft. For Load Case 2

bw 10 ft
= = 0.48  3
he 20.93
Therefore, Equation (6.10-5a) applies as follows:

1
Fd = × 2.2 ×1.0 × 2.0 ×10(20.93× 37.922 ) / 1, 000 = 662 kips.
2

This load is applied to the wall as a uniformly distributed pressure of 662/(10 × 20.93) =
3,163 psf over the lower 20.93 ft of the wall.
It is possible that the inundation occurs as a series of bores each with height less than
for LC2. In this case, the critical bore height would be

he 1 bw 10 ft
= → h=
e = = 3.33 ft
bw 3 3 3
because this would require consideration of Equation (6.10-5b). For a flow depth of 3.33 ft,
the flow velocity can be obtained from ASCE 7-16, Figure 6.8-1 as shown in Figure GA-38. The
resulting velocity is 17.1 ft/s. The bore loading is therefore computed for he = 3.33 ft and u =
17.1 ft/s.
For the 10 ft wide stairwell wall,

3
Fd = × 2.2 ×1.0 × 2.0 ×10(3.33×17.12 ) / 1, 000 = 32.1 kips.
4

This load is applied to the walls as a uniformly distributed pressure of 32.1/(10 × 3.33)
= 963 psf over the lower 3.33 ft of the walls. This will not govern when compared with the
3,183 psf pressure from Equation (6.10-5a).

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Appendix. Seaside, Oregon: Full Design Example 285
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

Figure GA-38. Determining u for Equation (6.10-5b) using ASCE 7-16, Figure 6.8-1.

A.13.3  Hydrodynamic Pressures Associated with


Slabs: Section 6.10.3
A.13.3.1  Flow Stagnation Pressure: Section 6.10.3.1
The mechanical room on Gridline D between Gridlines 5 and 6 is enclosed on all sides by
structural walls. Tsunami flow entering through the two door openings will result in flow
stagnation pressurization of this room, given by

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


286 Appendix. Seaside, Oregon: Full Design Example

1
Pp   s ITSU u 2 (6.10-8)
2
Assuming that the door openings are 7 ft high, the stagnation pressurization is based on
the maximum flow velocity occurring at this or greater depths, that is, when the door open-
ing is fully submerged. The flow velocity will therefore be the maximum of 37.92 ft/s which
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

occurs when the flow depth is 20.93 ft (Figure 6.8-1, LC2). Therefore,

1
Pp   2.2 1.0  37.922  1, 582 psf
2
The structural walls surrounding this room must be evaluated for an outward pressure
of 1,582 psf, in combination with gravity loads using the load combinations in Section 6.8.3.3.
The floor slab above this room must be designed for a net uplift pressure given by 0.9D +
FTSU = –0.9 × 100 + 1,582 = 1,492 psf upward. This will require additional top reinforcement in
this slab and possibly shear reinforcement around the slab perimeter. To reduce the amount
of additional reinforcement, one could perform a nonlinear analysis of the floor slab follow-
ing the provisions of ASCE 41. A simpler alternative may be to design the floor slab in the
mechanical room as a breakaway slab (Figure GA-39) to relieve pressure. This will apply to
all levels up to hmax.

Figure GA-39. Mechanical room breakaway floor panels applied to all levels up to hmax.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Appendix. Seaside, Oregon: Full Design Example 287

A.13.3.2  Hydrodynamic Surge Uplift at Horizontal Slabs:


Section 6.10.3.2
If slabs are submerged during the tsunami, they must be designed for uplift, with a specified
minimum of 20 psf (Section 6.10.3.2.1). The uplift may increase if the ground floor is sloped,
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

causing an upward component of flow velocity (Section 6.10.3.2.2). This is not the case for
this building.
The resulting minimum uplift of 20 psf is much smaller than the dead weight of the slab
(100 psf), and this uplift will not affect the slab design.

A.13.3.3  Tsunami Bore Flow Entrapped in Structural Wall-Slab


Recesses: Section 6.10.3.3
If a tsunami bore is entrapped in a structural wall-slab recess, then large pressures can
develop on the slab and wall (Section 6.10.3.3.1). Although tsunami bores are anticipated at
this location, the flow can pass freely around the wall elements in this building. Therefore,
this condition does not apply.

A.14  Debris Impact Loads: Section 6.11


The inundation depth at the site exceeds 3 ft. Therefore, exterior structural elements below
the flow depth must be designed for debris impact loads according to Section 6.11.

A.14.1  Alternative Simplified Debris Impact


Static Load: Section 6.11.1
In lieu of detailed debris impact analysis, the member can be designed for the maximum
static load given by

Fi = 330ITSUCo = 330 × 1.0 × 0.65 = 214.5 kips (6.11-1)

Because the building location is not within an impact zone for shipping containers, ships,
and barges, this force can be reduced by 50%, giving the limiting impact load of 107.25 kips.
As shown for the office building, this limiting impact load controls over the more detailed
debris impact calculations. This load must be applied to the 20 in. square exterior columns
as a static lateral load at points critical for flexure and shear, in combination with gravity
loads on the column. It is not combined with other tsunami loads and it need not be applied
to interior columns.
This equivalent static impact load of 107.25 kips must also be applied to any structural
walls on the perimeter of the building. This applies to the 28 ft wide elevator wall on the

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


288 Appendix. Seaside, Oregon: Full Design Example

ocean side and mechanical room wall on the inland side of the building (Gridlines A and D)
because impact must be considered during inflow and outflow conditions. Evaluation of the
wall capacity is based on a tributary wall width of half the wall height (Section 6.11.2). Because
the wall unbraced height is (12 ft – 8 in./12) = 11.33 ft, the tributary width is 5.67 ft.
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

A.15  Column Design for Tsunami Loads


A.15.1  Typical Exterior Column Design
A typical exterior column is chosen at Grid Intersection A-3 in Figure GA-13. The column
is not part of the lateral force–resisting system for seismic loads. It will be subject to the
same displacements as the lateral resisting system and needs to be detailed accordingly for
Seismic Design Category D. It is assumed to have a fixed base at the foundation; therefore,
a plastic hinge may form at the base of the column. The remainder of the column will not
form a plastic hinge, but the slab-column connections at each floor need to be detailed
appropriately for the seismic deformation. The 20 in. square column cross section shown in
Figure GA-40 was selected based on gravity load design and punching shear capacity of the
floor slabs.

Figure GA-40. Original exterior gravity column cross section at all floor levels based on Seismic Design
Category D design.

A.15.1.1  Column Gravity Loads


The gravity load tributary area for the typical exterior column is 28 by 14.58 ft. The dead load
at the base of the first floor column is PD = 406 kips and the reduced live load is PL = 54.2 kips.

A.15.1.2  Column Hydrodynamic Loads


Hydrodynamic loads from Load Case 2 will govern over LC1 and LC3 for this column. Analysis
of the column with the applied LC2 hydrodynamic loads results in the bending moment and
shear force diagrams as shown in Figure GA-41.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Appendix. Seaside, Oregon: Full Design Example 289
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

Figure GA-41. Hydrodynamic loading on exterior column of the Seaside residential building owing to
Load Case 2.

A.15.1.3  Column Debris Impact Loading


For debris impact loading, the static load of 107.25 kips is applied at all locations that will
cause maximum shear force and bending moments in the wall. A debris strike is assumed
to act just above and below each inundated floor slab for the maximum shear and near the
midheight of the clear wall height for maximum bending moments, up to hmax. Samples of the
resulting shear force and bending moment diagrams are provided in Figure GA-42 through
Figure GA-45. Similar diagrams and similar shear and bending moments would result if the
impact load was applied at the other end of each column and at all floor levels.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


290 Appendix. Seaside, Oregon: Full Design Example
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

Figure GA-42. Impact load applied at distance d away from the top of the column on the ground floor.

Figure GA-43. Impact load applied at the maximum water level on the fourth floor.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Appendix. Seaside, Oregon: Full Design Example 291
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

Figure GA-44. Impact load applied at midheight of the assumed lateral restraint points at the top and
bottom of the ground-floor column.

Figure GA-45. Impact load applied at midheight of the assumed lateral restraint points at the top and
bottom of the third-floor column.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


292 Appendix. Seaside, Oregon: Full Design Example

A.15.1.4  Column Load Combinations


Table GA-6 summarizes the maximum critical axial loads, bending moments, and shear
forces for the inundated levels of the exterior columns using the load combinations provided
in Section 6.8.3.3 both for hydrodynamic drag (hydro) and debris impact (impact). The origi-
nal column designs will be evaluated for these load combinations and modified if necessary.
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

Table GA-6. Load Combinations for the Seaside Residential Building Exterior Column

Moment Axial load Shear at d Shear at d + hc


(k-ft) (kips) (kips) (kips) Load combination

Floor 1

744 514 282 179 1.2D+FTSU+0.5L (hydro)

744 365 282 179 0.9D+ FTSU (hydro)

191 514 95 76 1.2D+ FTSU +0.5L (impact)

191 365 95 76 0.9D+ FTSU (impact)

Floor 2

615 441 217 113 1.2D+ FTSU +0.5L (hydro)

615 313 217 113 0.9D+ FTSU (hydro)

168 441 92 70 1.2D+ FTSU +0.5L (impact)

168 313 92 70 0.9D+ FTSU (impact)

Floor 3

175 367 24 24 1.2D+ FTSU +0.5L (hydro)

175 261 24 24 0.9D+ FTSU (hydro)

167 367 92 69 1.2D+ FTSU +0.5L (impact)

167 261 92 69 0.9D+ FTSU (impact)

Floor 4

41 294 6 6 1.2D+ FTSU +0.5L (hydro)

41 209 6 6 0.9D+ FTSU (hydro)

94 294 92 15 1.2D+ FTSU +0.5L (impact)

94 209 92 15 0.9D+ FTSU (impact)

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Appendix. Seaside, Oregon: Full Design Example 293

A.15.1.5  Exterior Column Design for Combined Gravity and


Tsunami Loads
The exterior column at Grid Intersection A-3 in Figure GA-13 will now be checked at all lev-
els for the load combinations in Table GA-6. Figure GA-46 through Figure GA-48 show the
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

interaction diagrams for the original exterior column along with the tsunami load combi-
nations. All three inundated levels of the column are able to resist the impact loads, but the
columns at the first two levels have to be strengthened to resist the bending moments from
tsunami hydrodynamic loads.

Figure GA-46. Interaction diagram for typical ground-floor exterior column showing tsunami load
combinations.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


294 Appendix. Seaside, Oregon: Full Design Example
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

Figure GA-47. Interaction diagram for typical second-floor exterior column showing tsunami load
combinations.

Figure GA-48. Interaction diagram for typical third floor-exterior column showing tsunami load
combinations.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Appendix. Seaside, Oregon: Full Design Example 295

A.15.1.6  New Typical Exterior Column Design


Based on the interaction diagrams shown in Figure GA-46 through Figure GA-48, the origi-
nal exterior columns are adequate for debris impact load, but the columns at the ground and
second floors must be strengthened to resist bending caused by the hydrodynamic loads.
Revised column designs were developed to satisfy the hydrodynamic loads as shown in Fig-
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

ure GA-49 and Figure GA-50. The interaction diagrams for these new columns are shown in
Figure GA-46 and Figure GA-47.

Figure GA-49. Exterior column cross section at the ground-floor level based on tsunami design
requirements.

Figure GA-50. Exterior column cross section at the second-floor level based on tsunami design
requirements.

A.15.1.7  Exterior Column Shear Design


For critical shears in the column at the ground floor, at the critical axial load combination
of (0.9D + FTSU) according to Equation (6.8-1a), Pu = 365 kips. The shear capacities of the 26 ×
26 in. columns with four leg #4 hoops at 4 in. on center in the end section and four leg #3
hoops at 5 in. on center in the center section are given by

ϕVn = ϕ(VC + VS)

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


296 Appendix. Seaside, Oregon: Full Design Example

 Pu   365, 000 


where VC = 2 f C ' 1 +  bd = 2 4, 000 1 +  26 × 23.295 / 1, 000 = 97 kips.
 2, 000 Ag   2, 000 × 26 × 26 

In the yield zone


Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

Av f y d  4  0.2   60, 000  23.295


VS = = = 280 kips
s 4 1, 000
In the rest of the column

Av f y d  3  0.2   60, 000  23.295


VS = = = 168 kips
s 5 1, 000
Therefore, in the end sections, ϕVn = 0.75 (97 + 280) = 283 kips, and in the center sec-
tions, ϕVn = 0.75 (97 + 168) = 199 kips.
At a distance d from the column end, Vu = 282 kips < φVn = 283 kips; therefore, the col-
umn is adequate for shear in the yield zones.
At a distance of d + hc from the column end, Vu = 179 kips < φVn = 199 kips; therefore, the
rest of the column is adequate for shear.
For critical shears in columns at the second floor, at the critical axial load combination
of (0.9D + FTSU) according to Equation (6.8-1a), Pu = 313 kips. The shear capacities of the 26 ×
26 in. columns with three leg #4 hoops at 4 in. on center in the end section and three leg #3
hoops at 5 in. on center in the center section are given by

ϕVn = ϕ(VC + VS)


 Pu   313, 000 
where VC = 2 f C ' 1 +  bd = 2 4, 000 1 +  26 × 23.436 / 1, 000 = 95 kips.
 2, 000 Ag   2, 000 × 26 × 26 

In the yield zone

Av f y d  3  0.2   60, 000  23.436


VS = = = 211 kips
s 4 1, 000
In the rest of the column

Av f y d  3  0.11  60, 000  23.436


VS = = = 93 kips
s 5 1, 000
Therefore, in the end sections, ϕVn = 0.75 (95 + 211) = 229 kips, and in the center sections,
ϕVn = 0.75 (95 + 93) = 141 kips.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Appendix. Seaside, Oregon: Full Design Example 297

At a distance d from the end of the column, Vu = 217 kips < φVn = 229 kips; therefore, the
column is adequate for shear at the yield zones.
At a distance d + hc from the end of the column, Vu = 113 kips < φVn = 141 kips; therefore,
the rest of the column is adequate for shear.
For critical shears in columns at the third floor, at the critical axial load combination of
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

(0.9D + FTSU) according to Equation (6.8-1a), Pu = 261 kips. The shear capacities of the original
20 × 20 in. columns with three leg #4 hoops at 4 in. on center in the yield zone and three leg
#3 hoops at 5 in. on center in the center section are given by

ϕVn = ϕ(VC + VS)


 Pu   
where VC = 2 f C ' 1 +
261, 000
 bd = 2 4, 000 1 +  20 ×17.56 / 1, 000 = 59 kips.
 2, 000 Ag   2, 000 × 20 × 20 

In the yield zone

Av f y d  3  0.2   60, 000 17.56


VS = = = 158 kips
s 4 1, 000
In the rest of the column

Av f y d  3  0.11  60, 000 17.56


VS = = = 70 kips
s 5 1, 000
Therefore, in the yield zone, ϕVn = 0.75 (59 + 158) = 163 kips, and in the rest of the column,
ϕVn = 0.75 (59 + 70) = 97 kips.
At a distance d from the end of the column, Vu = 92 kips < φVn = 163 kips; therefore, the
column is adequate for shear in the yield zones.
At a distance of d + hc from the end of the column, Vu = 69 kips < φVn = 97 kips; therefore,
the rest of the column is adequate for shear.

A.15.2  Typical Interior Column Design


A typical interior column is chosen at Grid Intersection B-3 in Figure GA-13. This gravity col-
umn is not part of the lateral force–resisting system for seismic loads. It will be subject to the
same displacements as the lateral resisting system and needs to be detailed accordingly for
Seismic Design Category D. It is assumed to have a fixed base at the foundation; therefore,
a plastic hinge may form at the base of the column. The remainder of the column will not
form a plastic hinge, but the slab-column connections need to be detailed appropriately for
the seismic deformation. The 20 in. square column cross section shown in Figure GA-51 was
selected based on gravity load design and punching shear capacity of the floor slabs.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


298 Appendix. Seaside, Oregon: Full Design Example

Similar to the interior column for the office building, this column is evaluated for the
hydrodynamic loads applied only to the column width, with no debris damming effect or
debris impact. The resulting load combinations are shown in Table GA-7. Figure GA-52
shows the interaction diagram for the original interior column at the ground-floor level
along with the tsunami load combinations. The original column is adequate to resist the
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

bending moments caused by hydrodynamic drag. In addition, the original column hoops are
adequate for the shear demand from the hydrodynamic drag. The interior columns therefore
require no modification to resist the tsunami loads.

Figure GA-51. Interior column cross section at all floor levels based on Seismic Design Category D design.

Table GA-7. Load Combinations for Seaside Residential Building Interior Column

Axial load Shear at d Shear at d + hc


Moment (k-ft) (kips) (kips) (kips) Load combination

Floor 1

63 812 24 15 1.2D+FTSU+0.5L (hydro)

63 566 24 15 0.9D+ FTSU (hydro)

Floor 2

52 677 18 10 1.2D+ FTSU +0.5L (hydro)

52 472 18 10 0.9D+ FTSU (hydro)

Floor 3

15 541 2 2 1.2D+ FTSU +0.5L (hydro)

15 377 2 2 0.9D+ FTSU (hydro)

Floor 4

3 406 0.5 0.5 1.2D+ FTSU +0.5L (hydro)

3 283 0.5 0.5 0.9D+ FTSU (hydro)

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Appendix. Seaside, Oregon: Full Design Example 299
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

Figure GA-52. Interaction diagram for ground-floor column of the residential building showing
tsunami load combinations.

A.15.3  Typical Exterior Wall Design


The exterior walls of the elevator core and mechanical room along Grid Lines A and D in
Figure GA-13 must be evaluated for hydrodynamic load and debris impact. The walls are part
of the lateral resisting system for seismic loads, acting as a shear wall for longitudinal forces
and boundary element for transverse forces. Seismic Design Category D design and detailing
of the 10 in. thick walls resulted in the reinforcement layout shown in Figure GA-53 through
Figure GA-55.
The exterior wall at the mechanical room is braced laterally at each level by the floor
slab, and the wall at the elevator core is braced at each level by a horizontal beam and trans-
verse beams in the plane of the floor slab. The exterior wall will therefore span vertically
between floor levels.
To compare the wall strength for both hydrodynamic and debris impact loads, both
analyses will be performed for an effective width of wall equal to one-half of the clear wall
height between floors. For the ground floor this width is h/2 = (12 ft – 8 in./12)/2 = 5.67 ft. The
critical shear force occurs at a distance d from the base of the wall, where d = 10 in. – 0.75 in.
– 1 in./2 = 8.75 in.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


300 Appendix. Seaside, Oregon: Full Design Example
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

Figure GA-53. Segment of exterior wall cross section at the ground-floor level based on Seismic Design
Category D design.

Figure GA-54. Segment of exterior wall cross section at the second-floor level based on Seismic Design
Category D design.

Figure GA-55. Segment of exterior wall cross section at the third- through seventh-floor levels based
on Seismic Design Category D design.

A.15.3.1  Shear Wall Gravity Loads


The elevator core wall has no floor loads, and the gravity load tributary width for the mechan-
ical room is 5.5 ft. The mechanical room wall will be used in this case for illustration. The
gravity loads at the ground level per foot width of wall are a dead load of 13.1 kips/ft and
reduced live load of 1.35 kips/ft.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Appendix. Seaside, Oregon: Full Design Example 301

A.15.3.2  Shear Wall Hydrodynamic Loads


Analysis of a 5.67 ft width of wall with the applied tsunami hydrodynamic pressure of 3,163 psf
from Load Case 2 results in a distributed load of 18 kips/ft up to the lower 20 ft − 11 in. of the
wall. The resulting bending moment and shear force diagrams are shown in Figure GA-56.
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

Figure GA-56. Hydrodynamic loading on 5.67 ft width of exterior wall of the Seaside residential
building owing to Load Case 2.

For debris impact loading, the equivalent static load of 107.25 kips resulting from a log
strike acts over an effective width of 5.67 ft at a point just below the slab at each inundated
floor for maximum shear and at the midheight of the clear wall height for maximum bending
moments. Sample shear force and bending moment diagrams for log impact at a distance,
d, from the end of the column are shown in Figure GA-57 and Figure GA-58. The resulting
shear force and bending moment diagrams for log impact at midheight of the column at
select floor levels are shown in Figure GA-59 and Figure GA-60. The resulting load combi-
nations for gravity, hydrodynamic, and impact loads on the 5.67 ft wide strip of exterior shear
wall are shown in Table GA-8.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


302 Appendix. Seaside, Oregon: Full Design Example
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

Figure GA-57. Impact load applied at distance d away from the top of the wall on the ground floor.

Figure GA-58. Impact load applied at distance d away from the bottom of the wall on the fourth floor.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Appendix. Seaside, Oregon: Full Design Example 303
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

Figure GA-59. Impact load applied at midheight of the assumed lateral restraint points at the top and
bottom of the ground-floor wall.

Figure GA-60. Impact load applied at midheight of the assumed lateral restraint points at the top and
bottom of the third-floor wall.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


304 Appendix. Seaside, Oregon: Full Design Example

Table GA-8. Results from Loading Conditions of the


Seaside Residential Building Exterior Shear Wall

Moment (k-ft) Axial load (kips) Shear at d (kips) Load combination

Floor 1
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

215 93 95 1.2D+FTSU+0.5L (hydro)

215 7 95 0.9D+ FTSU (hydro)

176 93 102 1.2D+ FTSU +0.5L (impact)

176 7 102 0.9D+ FTSU (impact)

Floor 2

178 80 76 1.2D+ FTSU +0.5L (hydro)

178 6 76 0.9D+ FTSU (hydro)

168 80 101 1.2D+ FTSU +0.5L (impact)

168 6 101 0.9D+ FTSU (impact)

Floor 3

51 66 7 1.2D+ FTSU +0.5L (hydro)

51 5 7 0.9D+ FTSU (hydro)

167 66 101 1.2D+ FTSU +0.5L (impact)

167 5 101 0.9D+ FTSU (impact)

Floor 4

12 53 2 1.2D+ FTSU +0.5L (hydro)

12 4 2 0.9D+ FTSU (hydro)

80 53 100 1.2D+ FTSU +0.5L (impact)

80 4 100 0.9D+ FTSU (impact)

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Appendix. Seaside, Oregon: Full Design Example 305

A.15.3.3  Existing Exterior Shear Wall Design for Combined


Gravity and Tsunami Loads
The shear wall chosen at Grid Line D in Figure GA-13 will now be checked at all levels for
combined flexure and axial loads caused by gravity and tsunami for load combinations
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

defined in Section 6.8.3.3. Figure GA-61 through Figure GA-64 show the interaction dia-
grams for the typical 5.67 ft width of exterior shear wall including the tsunami load combina-
tions from ground to fourth floors, respectively. The debris impact loads exceed the existing
wall bending capacity at the first three floor levels, and the hydrodynamic loads exceed the
wall bending capacity at the first two floor levels. The walls at the first three floor levels must
be strengthened to resist the tsunami loads, while the wall at the fourth level can remain
unchanged. The new wall reinforcement is shown in Figure GA-65 through Figure GA-67 for
the first three floors, respectively.

Figure GA-61. Interaction diagram for typical ground-floor exterior wall segment showing tsunami
load combinations.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


306 Appendix. Seaside, Oregon: Full Design Example
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

Figure GA-62. Interaction diagram for typical second-floor exterior wall segment showing tsunami
load combinations.

Figure GA-63. Interaction diagram for typical third-floor exterior wall segment showing tsunami load
combinations.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Appendix. Seaside, Oregon: Full Design Example 307
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

Figure GA-64. Interaction diagram for typical fourth-floor exterior wall segment showing tsunami load
combinations.

A.15.3.4  New Typical Shear Wall Design


The new walls are adequate for out-of-plane bending moments but must also be checked for
shear capacity. As shown in Section A.15.3.5, headed studs are required to increase the shear
capacity to resist the hydrodynamic and debris impact loads. Figure GA-65 through Fig-
ure GA-68 show the revised wall designs required to resist the tsunami loads. Figure GA-69
shows a cross section of the wall with shear stud rails included.

Figure GA-65. New exterior wall cross section at the ground-floor level based on tsunami design
requirements.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


308 Appendix. Seaside, Oregon: Full Design Example
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

Figure GA-66. New exterior wall cross section at the second-floor level based on tsunami design
requirements.

Figure GA-67. New exterior wall cross section at the third-floor level based on tsunami design
requirements.

Figure GA-68. New exterior wall cross section at the fourth-floor level based on tsunami design
requirements.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Appendix. Seaside, Oregon: Full Design Example 309
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

Figure GA-69. Stud rail diagram for Floors 1 through 4.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


310 Appendix. Seaside, Oregon: Full Design Example

A.15.3.5  Exterior Shear Wall Shear Design


For the shear wall at the ground level, the critical axial load according to Equation (6.8-1a)
(0.9D + FTSU) is Pu = 6.89 kips. Therefore, the axial load on a 5.67 ft (68 in.) wide strip of the
wall is
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

 P ×1, 000   
N u =  u × b =  6.89 ×1, 000 × 68 = 807 lb
 lw   580 

where lw = (28 ft + 11 ft +11 ft) × 12 – 2 × 10 in. (thickness) = 580 in.

The shear capacity of the existing shear wall is given by

ϕVn = ϕ(VC + VS)


 Nu   807 
where VC = 2 1 
  f 'c bd = 2 1   4, 000  68  8.75 / 1, 000 = 75 kips.
 2, 000 Ag   2, 000  68 10 

There is no shear reinforcement in the original wall, so ϕVn = ϕ(VC + VS) = 0.75 (75+0) =
56 kips.
Because VTSU = 102 kips > ϕVn = 56 kips, an additional shear strength of 61 kips is required.
Shear stud rails in the form of 3/8 in. diameter studs at 4 in. on center are added to the wall
at a horizontal spacing of 16 in. The shear capacity for the shear wall is then

ϕVn = ϕ(VC + VS)

where VC is 75 kips as before.

The shear force contributed by the headed studs is

n� �× Av� �× f y� �× d 4.25 × 0.11× 60 × 8.75


VS = = = 61 kips
s 4
where
Length 68 in.
n= = = 4.25 studs in a 68 in. width,
Rail spacing 16 in.
Av = 0.11 in2 for a 3/8 in. diameter stud,
smax = d/2 = 8.75/2 = 4.375 in., and
n ´ Av ´ f y ´ d 4.25´ 0.11´ 60 ´8.75
sneeded = = = 4.0 in.
Vs (needed) 61

Therefore, sused = 4 in.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Appendix. Seaside, Oregon: Full Design Example 311

Then, ϕVn = ϕ(VC + VS) = 0.75 (75 + 61) = 102 kips. So ϕVn = 102 kips = VTSU = 102 kips; there-
fore, the shear wall is now adequate for out-of-plane shear.
At midheight of the wall, VTSU = 63 kips > ϕVc = 0.75 × 75 = 56 kips; therefore, the stud rails
must extend up the wall to the maximum flow depth.
Similar calculations show that the same headed stud rails are adequate for the second
Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

through fourth floors.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Downloaded from ascelibrary.org by UNIVERSITY OF NEW SOUTH WALES on 10/01/20. Copyright ASCE. For personal use only; all rights reserved.

Tsunami Loads and Effects


@Seismicisolation
@Seismicisolation
Index
Downloaded from ascelibrary.org by 185.106.28.7 on 10/29/20. Copyright ASCE. For personal use only; all rights reserved.

Figures are indicated by f; tables are indicated by t.

acceptance criteria for structural performance, and, 179, 186–187, 187f; Load Case 1,
78–81, 79f 73–75, 74f, 95f, 102–105, 102f–103f, 106f,
Alaska, 14, 30–31 254f, 279; office building prototype,
Anak Krakatau, 14 252–253; reduction of effect, 94, 95f;
Applied Technology Council, 19 residential building prototype, 279; of
ASCE Tsunami Design Geodatabase: digital shipping containers, 157–158; trapped
elevation model resolution, 35; air and, 97, 99, 99f; uplift force, 93–95,
earthquake rupture unit source 99–101
functions, 67–68; EGLA, 43, 44f, 47–48,
48f, 56, 57f; geographic coverage of, California: applicability of design requirements,
225; Monterey, California project site, 30; Monterey prototype site. see
59f; NAVD88 and, 229; offshore wave Monterey, California, prototype site;
amplitude, 68; PHTA data, 32; prototype tsunami warning system, 14
site locations, 21, 22f; Seaside, Oregon Cascadia Subduction Zone, 14, 18, 21
project site, 240f; wave amplitude and Chile tsunami (2010), 1, 13, 170f
period data, 30–31, 30f; waveform
Chile tsunami (2015), 13
parameters, 65, 66f
Chock, Gary, 19–20, 21
asteroid strikes, 14, 225
closure coefficient: beam depth and, 117; closure
ratio and, 86; drag force and, 114, 116
barges as debris, 167–168, 173–174, 175f
closure ratio: load determination, 86–89; open
berms: facing systems, 201, 202f; lining failure,
structures, 87, 203; perforated walls,
199, 200f; plunging scour and, 197–198,
197f 131–132, 132f, 133f
boats as debris, 173–174, 174f column design: exterior columns, 263–275, 288–
boulders as submerged tumbling debris, 166–167, 297; interior columns, 275–278, 297–298,
167f, 176, 263 298f, 298t, 299f; for office buildings,
break-away walls and slabs, 95–97, 97f, 153–154, 263–279; for residential buildings,
154f, 204–205 288–311
bridges, 9, 11f, 179, 181f, 227 concrete as submerged tumbling debris, 166–167,
buildings. See specific types of buildings: 167f, 176, 263
reinforced concrete moment-resisting countermeasures: foundation design, 199–201,
frame office building prototype. see 200f, 202f; structural countermeasures.
office building prototype; reinforced see structural countermeasures
concrete shear wall residential building critical facilities: acceptance criteria, 80;
prototype. see residential building nonstructural component and system
prototype performance requirements, 217–218,
buoyancy: break-away walls and, 95–97, 97f; 219f; structural performance objectives,
buried utilities and, 227; foundations 71t, 73; TRC III designation, 38

313
@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


314 Index

debris accumulation: closure ratio and, 86–89, digital elevation models, 33, 35
86f; drag force and, 121–123; exclusion of displacements, 199
interior columns and walls, 229 drag coefficients, 114–115, 115t, 121, 121t
debris impact loads: barges, 173–174, 175f; drag force: on components, 116, 121–124, 121t,
duration of impact, 158, 160, 162; 260–261, 282–283; on exterior beam,
dynamic analysis, 174; effect of impacts 122–123; on exterior column, 122;
on structural materials, 158, 161; on interior column, 124; on overall
exclusion of interior columns and structure, 114–116, 118–120
walls, 228–229; exterior columns, 266,
267f–269f, 289, 290f–291f; floating earthquakes: foundations and, 90–91; near-
vehicles, 164, 174–175, 263; flowchart, source tsunamis and, 42, 79; seismic
Downloaded from ascelibrary.org by 185.106.28.7 on 10/29/20. Copyright ASCE. For personal use only; all rights reserved.

158, 159f; hazard assessment, 167–168, effects on structures, 90–91, 226;


170–171; Monterey, California office subsidence and, 42, 213; tsunami
building prototype site, 174–176; generation, 12–14, 67–68
physical modeling, 91; Seaside, Oregon energy grade line analysis (EGLA): amplified
prototype site, 261–263, 287–288; flow velocities, 62; directionality of
shipping containers, 157–158, 167–169, flow, 85; maximum inundation depth
169f–170f, 171–173, 173f; ships, 173–174, and flow velocities, 44–45, 45f, 51–52,
175f, 176f; simplified static load, 160–161; 52f–53f, 54, 55f; methodology, 46–58;
submerged tumbling debris, 95, 96f, overwashed island or peninsula, 56–58,
166–167, 167f, 176, 263; types of debris, 57f, 59f; performance requirement,
157; wood poles and logs. see wood pole 27; procedure, 47–54, 52f–53f, 54f–55f;
and log impacts Seaside, Oregon prototype site,
design provisions: applicability, 30–31; debris 244–245, 245f–248f, 248t; terrain
impact loads, 157–176. See also debris roughness, 58–60; theory of, 46–47, 46f;
impact loads; development of, 17–20, 18f; topographic transects, 43, 44f; TRC-
energy grade line analysis. See energy based requirements, 39–42, 42f; tsunami
grade line analysis; FAQs, 223–230; bores, 58, 61–62; vertical evacuation
flow characteristics, 39–42; foundation refuge structure design and, 210
design, 177–202. See also foundation energy head, 46–47
design; general requirements, 27–36; erosion, 182, 191, 191f, 192f
hydrodynamic loads, 111–156. See also evacuations, 14–17. See also vertical evacuation
hydrodynamic loads; hydrostatic refuge structures
loads, 93–110. See also hydrostatic evaluation criteria for structural performance:
loads; non-building structures, 221– acceptance criteria, 78–81, 79f;
222; nonstructural components and importance factors, 75–77, 76f, 77t; load
systems, 217–218, 219f; probabilistic cases, 73–75; load combinations, 77–78
tsunami hazard analysis, 32, 33f; exterior beam drag, 122–123
prototype buildings, 21–26. See also exterior columns: combined gravity and tsunami
prototype buildings; Seaside, Oregon loads, 271, 271f–272f, 293, 293f–294f;
prototype site example, 239–309. debris impact loads, 266, 267f–269f,
See also Seaside, Oregon, prototype 289, 290f–291f; drag, 122, 260, 282;
site; seismic design criteria, 25, 25t; gravity loads, 265, 271, 271f–272f, 288,
sequence, 27, 28f–29f, 30; structural 293, 293f–294f; hydrodynamic loads,
countermeasures, 203–206. See also 122, 260, 265, 266f, 282, 288, 289f; load
structural countermeasures; structural combinations, 269, 270t, 292, 292t; new
design procedures, 71–91. See also column design, 273, 273f, 295, 295f;
structural design procedures; symbols office building prototype, 263–275;
and notation, 34–36; terminology, residential building prototype, 288–297;
34–36, 35f; tsunami design zone maps, seismic design, 264, 265f; shear design,
32–34, 35; vertical evacuation refuge 273–275, 295–297; typical design, 263–
structures, 207–215. See also vertical 275, 264f, 288–297
evacuation refuge structures; wind exterior walls: break-away walls, 95–97;
speed and, 25, 25t combined gravity and tsunami

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Index 315

loads, 305, 305f–307f; design, 307, Fukushima Daiichi Nuclear Power Station, 217
307f–309f, 310–311; gravity loads,
300; hydrodynamic loads, 301, 301f, George E. Brown Jr. Network for Earthquake
302f–303f, 304t; new wall design, 307, Engineering Simulation (NEES), 19
307f–309f; nonstructural, 100–101, 100f; geotextile reinforcement systems, 201
typical design, 299–311, 300f gravity loads: exterior columns, 265, 271,
271f–272f, 288, 293, 293f–294f; exterior
facing systems for earth structures, 201, 202f walls, 300, 305, 305f–307f; interior
Federal Emergency Management Agency (FEMA), columns, 276, 277, 278f
19, 208 Great East Japan earthquake (2011), 20
Flood Insurance Rate Maps (FIRMs), 225 Guidelines for Design of Structures for Vertical
Downloaded from ascelibrary.org by 185.106.28.7 on 10/29/20. Copyright ASCE. For personal use only; all rights reserved.

flow stagnation pressure, 133, 135–137, 285–286, Evacuation from Tsunamis, P-646 (FEMA
286f 2008, 2012, 2019), 19, 208
flow velocity: amplification of, 62, 82–85, 228;
case loads, 75, 76f; directionality and, 85, Hawai’i: applicability of design requirements, 30;
120, 132–133, 134; EGLA, 44–45, 244–245, building code, 17, 20; Hilo prototype site.
245f–248f, 248t; inundation modeling, see Hilo, Hawai’i, prototype site; tsunami
63, 69; Load Case 2, 51–52, 53f, 54, 55f, warning system, 14; Waikiki prototype
254–255, 256f, 280, 281f; office building site. see Waikiki, Hawai’i, prototype site
prototype, 254–255, 256f; overwashed hazard assessment for debris impacts, 167–168,
islands or peninsulas, 58, 59f; physical 170–171
modeling, 91; residential building Hilo, Hawai’i, prototype site: debris impact
prototype, 280, 281f; Seaside, Oregon hazard assessment, 170–171, 172f;
prototype site, 244–245, 247f–248f,
description of, 21; design wind speed,
248t, 254–255, 256f, 280, 281f; upstream
25, 25t; seismic design, 25t; seismic
obstructing structures and, 83, 84f, 84t
design parameters, 25–26, 26t; tsunami
fluid density, 81–82
bore occurrence, 62
foundation design: countermeasures, 199–201;
Hurricane Sandy (2012), 217
displacements, 199; erosion and, 191,
hurricane storm surge loads, 111
191f, 192f; facing systems, 201, 202f;
hydrodynamic loads: detailed lateral forces,
failure mechanisms, 179; fill and, 199;
113–133; diagonal flow and, 120; drag
geotextiles, 201; ground improvement,
201; horizontal soil loads, 182, 183f, 198; coefficients, 114–115, 115t, 121, 121t; drag
hydrostatic surcharge pressure, 110, force on components, 116, 121–124, 121t,
110f; load and effect characterization, 260–261, 282–283; drag force on overall
180, 182, 183f, 184–199, 184f–185f; structure, 114–116, 118–120; exterior
loss of soil shear strength, 189–190; beam drag, 122–123; exterior columns,
performance-based criteria, 199; 122, 260, 265, 266f, 282, 288, 289f;
physical modeling, 91; piles, 187–188, exterior walls, 300, 301f, 302f–303f,
187f, 188f; plunging scour, 177, 178f, 179, 304t; flow stagnation pressure, 133, 135–
180f, 194, 196–199; reinforced earth 137, 285–286, 286f; interior columns, 124,
systems, 201; resistance factors for 260–261, 276, 276f, 282–283; Load Case
stability analysis, 179–180; scour and, 177, 1, 102–105, 102f–103f, 106f, 279; office
177f–178f, 179, 180f, 181f, 193–198; seismic building prototype, 252–253; perforated
effects of subduction zone earthquakes, walls and. See perforated walls; physical
90–91; slab-on-grade, 201; sustained modeling, 91; Seaside, Oregon prototype
flow scour, 177f, 184f–185f, 193, 193f–194f; site, 254f, 260–261, 279, 282–283;
uplift and seepage forces, 185–189, 186f, simplified equivalent static lateral force,
187f, 188f, 189f 112–113; on slabs, 133–156, 285–287; surge
Froude number: EGLA and, 44, 50–51; for uplift at horizontal slabs, 134, 138–139,
overwashed islands or peninsulas, 56, 140, 287; tsunami bores. See tsunami
58; scour and, 193–194, 195f; for tsunami bores; tsunamis vs. hurricanes, 111; walls
bores, 58, 124; wall-slab recess pressure angled to flow, 132–133; wave breaking
and, 124 and, 111–112

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


316 Index

hydrostatic loads: buoyancy, 93–105. See also loads; hydrostatic loads, 93–110. see also
buoyancy; residual water surcharge, hydrostatic loads
108–110, 108f, 110f; surcharge pressure Long Beach Peninsula, Washington, 56, 57f, 58,
on foundations, 110, 110f; unbalanced 59f
lateral force, 105–107, 106f, 107f
Manning’s coefficient, 50–54, 52f–55f, 58, 60
Indian Ocean tsunami (2004), 1, 13, 19, 143, 145f maps and mapping. See also ASCE Tsunami
Indonesia, 13–14 Design Geodatabase: evacuation maps,
interior columns: combined gravity and tsunami 225; tsunami design zones, 31–35, 225
loads, 277, 278f; drag, 124, 260–261, 282– masonry structures. see reinforced concrete and
283; gravity loads, 276; hydrodynamic masonry structures
Downloaded from ascelibrary.org by 185.106.28.7 on 10/29/20. Copyright ASCE. For personal use only; all rights reserved.

loads, 276, 276f; load combinations, maximum considered tsunami (MCT): design
277, 277t, 298, 298t; office building provisions, 30–31, 30f, 226; vertical
prototype, 275–278, 276f; typical design, evacuation refuge structure design, 7
275–278, 276f, 297–298, 298f, 298t, 299f mean high water (MHW), 48, 229
International Building Code (IBC), 20, 224 meteor strikes, 14, 225
inundation: build environment and, 228; case Minimum Design Loads and Associated Criteria
loads, 75, 76f; earthquake rupture unit for Buildings and Other Structures
source functions, 67–68; evacuation (ASCE 7-16): debris impact loads.
of inundation zone, 15; flow analysis, see debris impact loads; definitions,
39–42, 42f; flow velocity. see flow symbols, and notation, 34–36,
velocity; limit defined, 35; maximum 35f; design provisions. see design
provisions; development of Chapter 6,
depth, 44–45, 45f, 69; maximum
17–20; energy grade line analysis. See
momentum flux, 69–70; offshore
energy grade line analysis; evacuation
wave amplitude, 68; procedures for
structures. see vertical evacuation
modeling, 68–70; PTHA and, 63, 67;
refuge structures; foundation design.
representative design parameters, 69;
see foundation design; hydrodynamic
runup and, 43–62; sea level change
loads. See hydrodynamic loads;
and, 42; Seaside, Oregon prototype
hydrostatic loads. See hydrostatic
site, 244–245, 245f–246f, 255–257, 257f,
loads; inundation. See inundation;
280; site-specific modeling, 63–70,
nonstructural components and systems.
85; tsunami design zone maps and, See nonstructural components and
32–33; tsunamigenic sources and, systems; probabilistic tsunami hazard
67; uncertainties in modeling, 68; analysis. See probabilistic tsunami
waveforms, 64–65, 66f–67f hazard analysis; prototype buildings.
island flank failure, 14, 225 see prototype buildings; prototype sites.
island washover, 56, 57f, 58, 59f See specific project sites; risk categories.
See Tsunami Risk Categories; scope of,
Japan. See also Tohoku tsunami: warning and 30–31; structural countermeasures. see
evacuation plans, 15, 16f structural countermeasures; structural
Java, 14 design procedures. see structural design
procedures
landslides, 13–14, 31 momentum flux, 51–52, 54, 54f, 55f, 69–70
lateral force-resisting system evaluation, 78–80, Monterey, California, prototype site: debris
79f, 258, 259f, 281–282, 281f impact load, 174–176; description of,
life safety level, 78 21; design wind speed, 25, 25t; energy
Lituya Bay tsunami (1958), 13 grade line analysis, 47–54, 48f, 52f–53f,
load combinations: description of, 77–78; on 54f–55f; scour depth, 194, 195f; seismic
exterior columns, 269, 270t, 292, 292t; design, 25–26, 25t, 26t; tsunami bore
on interior columns, 277, 277t, 298, 298t occurrence, 59f, 62; waveforms, 65,
loads: debris impact loads, 157–176. see also 66f–67f
debris impact loads; hydrodynamic
loads, 111–156. See also hydrodynamic National Crash Analysis Center, 164

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Index 317

National Oceanic and Atmospheric plunging scour: cause of, 194; description of,
Administration (NOAA), 17, 19 177, 178f, 180f; foundations and, 179;
National Tsunami Hazard Mitigation Program over berm, 197–198, 197f; over seawall,
(NTHMP), 17–19 196–197, 196f
National Tsunami Warning Center, 14 Preuss, Jane, 18
Network for Earthquake Engineering Simulation probabilistic tsunami hazard analysis (PTHA):
Research (NEESR), 161 description of, 32, 33f; events
non-building structures, 221–222 considered, 225; inundation modeling
nonstructural components and systems, 100–101, and, 63, 67; offshore wave amplitudes
100f, 217–218, 219f and, 229–230; uncertainties and, 68;
North American Vertical Datum (NAVD88), 48, vertical evacuation refuge structure
Downloaded from ascelibrary.org by 185.106.28.7 on 10/29/20. Copyright ASCE. For personal use only; all rights reserved.

229 design and, 210–211, 213


notation, 34–36 progressive collapse prevention, 81, 173–174
numerical models, 33–34, 85 prototype buildings: development of, 21; locations,
21, 22f; reinforced concrete moment-
office building prototype: debris impact loads, resisting frame office building. See office
174–176, 261–263; description of, 23, 23f; building prototype; reinforced concrete
drag force on components, 260–261; shear wall residential building. See
exterior columns, 260, 263–275. See also residential building prototype; seismic
exterior columns; exterior nonstructural design, 25–26, 26t; seismic design
wall failure load, 100–101, 100f; interior criteria, 25t; wind speed at, 25, 25t
columns, 260–261, 275–278, 276f. See Puget Sound, 14
also interior columns; lateral force-
resisting system evaluation, 258, 259f; reinforced concrete and masonry structures:
Load Case 1, 252–253; Load Case 2, moment-resisting frame office building
254–255, 256f; Load Case 3, 256–257, prototype. See office building prototype;
257f, 280; as open structure, 87; Seaside, resiliency of, 4, 5f, 7–9, 8f, 9f; seismic
Oregon project site, 248–249, 249f, design parameters for prototypes,
252–278; seismic design, 25–26, 25t, 26t; 25–26, 26t; shear wall residential
simplified debris impact static load, 261; building prototype. See residential
submerged tumbling debris impact, 263; building prototype; Tohoku tsunami
vehicle impact, 263; wood log and pole damage, 1, 3f–4f
impact, 262–263 reinforced earth systems, 201
open structures, 87, 203–204, 204f reliability analysis, 75–77, 76f, 77t
Oregon: applicability of design requirements, residential building prototype: bore load, 129;
30; Seaside prototype site. See Seaside, closure coefficient, 88–89; debris
Oregon prototype site; tsunami warning impact loads, 287–288; description
system, 14 of, 24, 24f; drag force on components,
overturning. See buoyancy 282–283; exterior columns, 282,
overwashed islands and peninsulas, 56, 57f, 58, 288–297; exterior walls, 299–311, 300f;
59f flow stagnation pressure, 137, 285–286,
286f; gravity load, 300; hydrodynamic
Pacific Tsunami Warning Center, 14 drag, 88–89, 282–283; hydrodynamic
Palu-Koro Fault, 14 pressures on slabs, 285–287; interior
Palu tsunami (2018), 14 columns, 282–283, 297–298, 298f, 298t,
peer review requirements, 81, 199, 214–215 299f; lateral force-resisting system
peninsula washover, 56, 57f, 58, 59f evaluation, 281–282, 281f; Load Case
perforated walls: closure ratio, 131–132, 132f, 133f; 1, 279; Load Case 2, 280, 281f; as open
uplift and, 149, 150f, 152, 153 structure, 87; Seaside, Oregon project
performance requirements: nonstructural site, 249–250, 250f, 279–311; seismic
components and systems, 217–218, 219f; design, 25–26, 25t, 26t; simplified debris
structural components, 80–81 impact static load, 287–288; slab over
physical modeling, 91 sloping grade uplift, 140; structural
pile foundation design, 187–188, 187f, 188f walls, 283–284, 285f; surge uplift, 287

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


318 Index

residual water surcharge load, 108–110, 108f, 110f seepage forces on foundations, 185–189, 186f,
resistance factors for stability analysis, 179–180 187f, 188f, 189f
retrofits, 227 seismic design: columns, 264, 264f, 265f;
Rosedale Fault, 31 prototype buildings, 25–26, 25t, 26t
runup: defined, 35; EGLA, 43–52; evaluation shear design: exterior columns, 273–275,
procedure, 39; flow analysis, 39–42, 295–297; exterior walls, 307, 307f–309f,
42f; Seaside, Oregon project site, 239, 310–311; interior columns, 277–278
240t; terrain roughness and, 58; tsunami shear strength of soil, 189–190
design zone maps and, 32–33 shear walls. See exterior walls
shipping containers: buoyancy of, 158; as debris,
Samoa tsunami (2009): fatalities, 1; inundation 157–158; simplified static load analysis,
Downloaded from ascelibrary.org by 185.106.28.7 on 10/29/20. Copyright ASCE. For personal use only; all rights reserved.

model, 210, 211f; tumbling debris, 95, 96f; 160–161; site hazard assessment, 167–
wharf uplift, 143, 144f, 146f 168, 169f–170f, 170–171
scour: countermeasures, 199–201, 200f, 202f; ships as debris, 160, 167–168, 175f, 176f
depth, 193–194, 194f, 195f; erosion signage, 15, 16f, 17f
vs., 182; flow cycles considered, 90; simplified debris impact static load, 160–161, 261,
inundation modeling and, 69; plunging 287–288
scour, 177, 178f, 179, 180f, 194, 196–199; simplified equivalent static lateral load, 112–113,
slab-on-grade and, 201, 229; sustained 252, 279
flow scour, 177f, 184f–185f, 193, 193f–194f slab-on-grade construction: buoyancy
scour enhancement parameter, 190 reduction, 74, 94–95, 95f; foundation
sea level change, 42, 241 design, 201, 229; tumbling debris and,
Seaside, Oregon prototype site: assumed 95, 96f; uplift pressure, 139, 140, 141f
conditions, 251–252; column design, slabs: bore flow entrapment, 139–156, 287;
263–278, 288–311; debris impact loads, flow stagnation pressure, 133, 135–137,
261–263, 266, 267f–269f, 287–288, 289, 285–286, 286f; hydrodynamic pressures
290f–291f; design wind speed, 25, on, 133–156, 285–287; uplift, 134, 138–139,
25t; EGLA, 244–245, 245f–248f, 248t; 140, 141f, 287
exterior column design, 263–275, soils. See foundation design
288–297. see also exterior columns; soliton fission, 68
full design, 239–309; interior column strike-slip earthquakes, 14
design, 275–278, 297–298, 298f, 298t, structural countermeasures: break-away wall
299f. see also interior columns; lateral design, 204–205; open structures, 203–
force-resisting system evaluation, 258, 204, 204f; tsunami barriers, 204, 206
259f, 281–282, 281f; Load Case 1, 252– structural design procedures: acceptance
253, 254f, 279; Load Case 2, 254–255, criteria, 78–81, 79f; closure ratio, 86–89;
256f, 280, 281f; Load Case 3, 255–257, directionality of flow, 85, 120, 132–133,
257f, 280; load summary, 250, 251t; 134; flow cycles considered, 90; flow
location, 21, 22f; MCT data, 30–31, 30f; velocity amplification, 82–85; fluid
prototype buildings, 248–250, 249f, 250f; density, 81–82; importance factors,
reinforced concrete moment-resisting 75–77, 77t; load cases, 73–75, 76f; load
frame office building, 23, 23f, 248–249, combinations, 77–78; performance
249f, 252–278; reinforced concrete shear evaluation criteria, 73–81; performance
wall residential building, 24, 24f, 249– objectives, 71–73, 71t; physical modeling,
250, 250f, 279–282; residential building, 91; seismic effects on foundations,
279–311; sea level change, 241; seismic 90–91
design, 25–26, 25t, 26t; site description, structural fill, 199
239, 240f, 240t; topographic profiles, 241, structural steel buildings: applicability of design
242f–243f; tsunami bore occurrence, 62, provisions, 21; buoyancy, 93; foundation
243–244, 244f failure, 189, 189f; resiliency of, 4, 5f, 7–8,
Seattle Fault, 31 8f; Tohoku tsunami damage, 1, 4f, 10f;
Seattle-Tacoma region, 14 vertical evacuation refuge structures,
seawalls: lining failure, 199, 200f; plunging scour 207, 208f
and, 179, 180f, 196–197, 196f structural walls, 283–284, 285f

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


Index 319

subduction zone earthquakes, 12–13, 12f objectives, 71t, 72–73; seismic effects on
submarine landslides, 13–14 foundations, 90–91
Sulawesi, 14 TRC IV structures: acceptance criteria, 80;
Sumatra, 14 applicability of design requirements,
Sunda Strait tsunami (2018), 13–14 223; design requirements, 31, 37;
surf similarity, 39–41 flow analysis, 41–42, 42t; importance
sustained flow scour, 177f, 184f–185f, 193, factors, 75–77, 77t; inundation modeling
193f–194f requirement, 63; non-building
symbols, 34–36 structures, 221–222; performance
objectives, 71t, 73
Tacoma Fault, 31 tsunami barriers. See also berms; seawalls: non-
Downloaded from ascelibrary.org by 185.106.28.7 on 10/29/20. Copyright ASCE. For personal use only; all rights reserved.

terminology, 34–36, 35f building structure design requirements,


terrain roughness, 58–60, 63. See also Manning’s 221–222; as structural countermeasure,
coefficient 204, 206
timber-framed residential structures, 1, 2f–3f, 97, tsunami bores: angled walls and, 134; definition
97f, 223 of, 226; EGLA, 58, 61–62; entrapment,
Tohoku tsunami (2011): ASCE damage survey, 139–156, 141f, 142f, 143f, 287; flow velocity
20; bore strike, 127, 128f; buoyancy and, amplification, 82–85; inundation
74f, 75, 93, 94f, 97, 97f; critical facility modeling, 68–69; perforated walls and,
damage, 217–218, 219f; damage caused 131–132, 132f, 133f, 149, 150f, 152–153;
by, 1, 2f–11f, 4–9; debris damming, Seaside, Oregon project site, 243–244,
86, 86f; erosion and, 191, 191f, 192f; 244f; uplift pressure reduction, 147,
149, 150f, 151–156, 152f, 154f; vertical
fatalities, 15; fishing vessels as debris,
structural components and, 124–132,
174f; flow stagnation, 135–136, 135f,
125f, 126f–127f, 128f; wall-slab recess
136f; generation of, 13; residual water
pressure load, 144, 146–147, 147f, 148–149,
surcharge load, 108, 108f; scour,
150f; wave flume experiments, 139–143,
177f–178f; shipping container yard
141f, 142f, 143f; wharf uplift, 143–144, 144f,
damage, 169f; ships as debris, 176f;
145f–146f
soliton fission and, 68; uplift and
Tsunami Design Geodatabase. See ASCE Tsunami
seepage pressure, 186–189, 186f, 187f,
Design Geodatabase
188f, 189f; vertical evacuation refuge
Tsunami Hazard Tool. See ASCE Tsunami Design
structures, 207–208, 208f, 209f, 210; Geodatabase
wharf uplift, 144 Tsunami Loads and Effects Subcommittee
topographic transects: EGLA and, 43, 44f, 47–49, (TLESC), 17, 20
48f, 56, 57f; Seaside, Oregon project site, Tsunami Risk Categories (TRC): description of,
239, 240f, 241, 242f–243f 37–38; flow characteristics and, 39–42;
TRC. See Tsunami Risk Categories TRC I structures, 38; TRC II structures.
TRC I structures, 38 See TRC II structures; TRC III
TRC II structures: acceptance criteria, structures. see TRC III structures; TRC
78–79, 79f; design requirements, IV structures. See TRC IV structures
16–17, 31, 38; enactment of design tsunamis. See also specific tsunamis: damage
requirements, 224; exclusion from from, 1–9, 2f–11f; debris impact loads,
design requirements, 223; flow analysis, 157–176. See also debris impact loads;
39–40, 42t; importance factors, 75–77, design provisions. See design provisions;
77t; performance objectives, 71t, 72; evacuation, 14–17, 207–215. See also
prototype buildings, 21; seismic effects vertical evacuation refuge structures;
on foundations, 90–91 FAQs, 223–230; flow characteristics,
TRC III structures: acceptance criteria, 39–42; foundation design, 177–202. See
78–79, 79f; applicability of design also foundation design; generation
requirements, 223; design requirements, mechanisms, 12–14, 67–68; historical,
31, 37–38; flow analysis, 39–40, 42t; 1–9; hydrodynamic loads, 111–156. See
importance factors, 75–77, 77t; non- also hydrodynamic loads; hydrostatic
building structures, 221; performance loads, 93–110. See also hydrostatic loads;

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


320 Index

tsunamis (continued) vertical evacuation refuge structures:


inundation. see inundation; load construction documents, 214;
summary, Seaside, Oregon prototype description of, 15–17, 16f–17f; evolution of
site, 250, 251t; non-building structure design, 207–208, 210; laydown impacts,
design requirements, 221–222; 214; live load, 212; minimum elevation
nonstructural component and requirements, 210–213, 211f, 212f, 213f;
system performance requirements, open structures, 204f; peer review of
217–218, 219f; prototype buildings, design, 214–215; Tohoku tsunami and,
21–26. see also prototype buildings; 4–6, 5f, 6f, 7f
risk categories. See Tsunami Risk
Categories; Seaside, Oregon prototype Waikiki, Hawai’i, prototype site: design wind
Downloaded from ascelibrary.org by 185.106.28.7 on 10/29/20. Copyright ASCE. For personal use only; all rights reserved.

site design example, 237–309. See speed, 25, 25t; location, 21, 22f;
also Seaside, Oregon prototype site; reinforced concrete moment-resisting
structural countermeasures, 203–206. frame office building, 23; reinforced
See also structural countermeasures; concrete shear wall residential building,
structural design procedures, 71–91. 24; seismic design, 25–26, 25t, 26t;
See also structural design procedures; tsunami bore occurrence, 62
warning and evacuation, 14–17; wave walls: exterior. See exterior walls; nonstructural,
propagation, 13 100–101, 100f; structural, 283–284, 285f
tumbling boulders and concrete, 263 warning systems, 14–15
Washington: applicability of design
unbalanced lateral hydrostatic force, 105–107, requirements, 30; Long Beach Peninsula
106f, 107f washover, 56, 57f, 58, 59f; tsunami
unreinforced/lightly reinforced masonry warning system, 14
structures, 1, 3f wastewater treatment facilities. See critical
uplift: break-away walls and, 154; buoyancy facilities
effects. see buoyancy; foundations waveforms: inundation modeling and, 64–65,
and, 179, 185–189, 186f, 187f, 188f, 189f; 66f–67f; offshore wave amplitude, 68,
horizontal slabs and, 134, 138–139, 140; 229–230; wave breaking and, 111–112
inundation depth and, 147, 149, 151–152; wharf uplift, 9, 143–144, 144f, 145f–146f
perforated walls and, 149, 150f, 152, 153; windows, 97–99, 101, 103
slab openings and, 152–153, 154f, 155–156, wood pole and log impacts: with column, 165–
287; Tohoku tsunami and, 9, 10f–11f 166; debris, 157, 160–164; duration, 161,
US Geological Survey (USGS), 32 162f; Monterey, California project site,
174–175; Seaside, Oregon prototype site,
vehicles as floating debris, 164, 174–175, 263 262–263; shipping containers and, 163,
Vertical Evacuation from Tsunamis: A Guide for 163f; utility pole failure, 214
Community Officials, P-646A (FEMA
2009), 19 Yeh, Harry, 18

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


About the Author
Downloaded from ascelibrary.org by 185.106.28.7 on 10/29/20. Copyright ASCE. For personal use only; all rights reserved.

Ian N. Robertson, Ph.D., S.E., is the Arthur N. L. Chiu Distin-


guished Professor of structural engineering in the Department
of Civil and Environmental Engineering at the University of
Hawai‘i at Mānoa. He received his Bachelor of Science degree
in civil engineering at the University of the Witwatersrand in
Johannesburg, South Africa, and his master’s and doctorate
degrees from Rice University, Houston, Texas. After receiving
his doctorate, he spent three years with Walter P Moore and
Associates, a structural engineering consulting company in
Houston, before being licensed as a professional engineer in
the state of Texas. In 1992 he accepted a faculty position at
the University of Hawai‘i at Mānoa where he teaches struc-
tural engineering courses on Structural Loading, Reinforced
and Prestressed Concrete, and Structural Steel Design. He is a registered structural engi-
neer in the state of Hawai‘i and is a past president of the Structural Engineers’ Association
of Hawai‘i.
His research interests include the performance of steel and concrete structures during
tsunami, earthquake, hurricane, and other extreme loading events, and the long-term behav-
ior and durability of reinforced and prestressed concrete and steel structures. Robertson
worked with the Applied Technology Council to develop both the first and second editions
of FEMA P-646, Guidelines for Design of Structures for Vertical Evacuation from Tsunamis,
and recently edited the third edition released in 2019. He has also led or participated in
post-tsunami reconnaissance surveys after the 2009 Samoa, 2010 Chile, 2011 Japan Tohoku,
and 2018 Palu tsunamis. From 2005 to 2010, he directed a major NSF-funded NEES research
project to develop Performance-Based Tsunami Engineering, PBTE. This project led to cre-
ation of a new subcommittee of ASCE 7 that subsequently developed the first comprehen-
sive probabilistic tsunami design provisions for coastal buildings and other structures in
the United States. Robertson serves as vice chair of this ASCE 7 Tsunami Loads and Effects

321
@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects


322 Tsunami Loads and Effects

subcommittee. He is the author of dozens of research articles on the effects of tsunamis and
storm surges, as well as coauthor of Tohoku, Japan, Earthquake and Tsunami of 2011: Perfor-
mance of Structures under Tsunami Loads. He has made numerous presentations on tsunami
design and community preparedness at workshops, conferences and webinars around the
world.
Downloaded from ascelibrary.org by 185.106.28.7 on 10/29/20. Copyright ASCE. For personal use only; all rights reserved.

@Seismicisolation
@Seismicisolation

Tsunami Loads and Effects

You might also like