You are on page 1of 98

https://www.tarjomano.

com/order

DESIGN AND
PERFORMANCE OF
TALL BUILDINGS
FOR WIND
ASCE MANUALS
AND REPORTS ON
ENGINEERING PRACTICE
NO. 143

EDITED BY PREETAM BISWAS


JOHN PERONTO

TASK COMMITTEE FOR THE


DESIGN AND PERFORMANCE OF
TALL BUILDINGS FOR WIND
ASCE Manuals and Reports on Engineering Practice No. 143

Design and
Performance of Tall
Buildings for Wind

Edited by Preetam Biswas and John Peronto

Prepared by
Task Committee for the Design and Performance of Tall Buildings for
Wind of the Structural Engineering Institute of the
American Society of Civil Engineers

Published by the American Society of Civil Engineers


Library of Congress Cataloging-in-Publication Data

Names: Biswas, Preetam, editor. | Peronto, John, editor. | Task Committee for the Design and
Performance of Tall Buildings in Wind author.
Title: Design and performance of tall buildings for wind / prepared by Task Committee for the
Design and Performance of Tall Buildings in Wind; edited by Preetam Biswas, P.E., and John
(Sp.) - Peronto.
Description: Reston, Virginia : American Society of Civil Engineers, [2020] | Series: ASCE
manuals and reports on engineering practice ; no. 143 | Includes bibliographical references
and index. | Summary: “Design and Performance of Tall Buildings for Wind, MOP 143
provides a framework for the design of tall buildings for wind, based on the current state-
of-practice in tall building structural design and wind tunnel testing”– Provided by publisher.
Identifiers: LCCN 2020033269 | ISBN 9780784415658 (paperback) | ISBN 9780784483121
(adobe pdf)
Subjects: LCSH: Tall buildings–Aerodynamics. | Wind resistant design.
Classification: LCC TH891 .D37 2020 | DDC 690/.21–dc23
LC record available at https://lccn.loc.gov/2020033269

Published by American Society of Civil Engineers


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

Any statements expressed in these materials are those of the individual authors and do not
necessarily represent 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, recommendation, 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,
statutes, 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 materials 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://ascelibrary.org) and using the “Permissions” link.

Errata: Errata, if any, can be found at http://dx.doi.org/10.1061/9780784415658.

Copyright © 2020 by the American Society of Civil Engineers.


All Rights Reserved.
ISBN 978-0-7844-1565-8 (soft cover)
ISBN 978-0-7844-8312-1 (PDF)

Manufactured in the United States of America.


25 24 23 22 21 20 1 2 3 4 5
MANUALS AND REPORTS ON
ENGINEERING PRACTICE

(As developed by the ASCE Technical Procedures Committee,


July 1930, and revised March 1935, February 1962, and April 1982)

A manual or report in this series consists of an orderly presentation of


facts on a particular subject, supplemented by an analysis of limitations and
applications of these facts. It contains information useful to the average
engineer in his or her everyday work, rather than findings that may be use-
ful only occasionally or rarely. It is not in any sense a “standard,” however,
nor is it so elementary or so conclusive as to provide a “rule of thumb” for
nonengineers.

Furthermore, material in this series, in distinction from a paper (which


expresses only one person’s observations or opinions), is the work of a com-
mittee or group selected to assemble and express information on a specific
topic. As often as practicable the committee is under the direction of one or
more of the Technical Divisions and Councils, and the product evolved has
been subjected to review by the Executive Committee of the Division or
Council. As a step in the process of this review, proposed manuscripts
are often brought before the members of the Technical Divisions and
Councils for comment, which may serve as the basis for improvement.
When published, each manual shows the names of the committees by
which it was compiled and indicates clearly the several processes through
which it has passed in review, so that its merit may be definitely
understood.

In February 1962 (and revised in April 1982), the Board of Direction


voted to establish a series titled “Manuals and Reports on Engineering
Practice,” to include the manuals published and authorized to date, future
Manuals of Professional Practice, and Reports on Engineering Practice. All
such manual or report material of the Society would have been refereed in a
manner approved by the Board Committee on Publications and would be
bound, with applicable discussion, in books similar to past manuals.
Numbering would be consecutive and would be a continuation of present
manual numbers. In some cases of joint committee reports, bypassing of
journal publications may be authorized.

A list of available Manuals of Practice can be found at https://www.asce.org/


bookstore.
AUTHORS

Preetam Biswas, Skidmore, Owings & Merrill


Preetam Biswas is a director of structural engineering at Skidmore, Owings
& Merrill (SOM). He has led the design of multiple tall buildings, airports,
stadiums, and convention centers around the globe. He has authored many
technical papers focusing on tall buildings and other structural system
innovations and served as the chair for the task committee that authored
this Manual of Practice.

John Peronto, Thornton Tomasetti, Inc.


John Peronto is a senior principal at Thornton Tomasetti and has led the
design of many tall-to-megatall buildings, as well as unique and iconic struc-
tures around the world. He is highly published and a leader in professional
organizations such as ASCE, ACI, IStructE, ICE, CTBUH, CCHRB, and has
served as the chair for the Tall Buildings Committee of SEI since 2016.

Kevin Aswegan, Magnusson Klemencic Associates


Kevin Aswegan is an associate with Magnusson Klemencic Associates
(MKA) in Seattle, where his design experience focuses on tall buildings
in areas of high wind and seismicity in the United States and around the
world. He is actively involved in wind-related research, has authored many
articles on performance-based wind design, and is an active member of the
Standard ASCE 7 update process.

Roy Denoon, CPP Wind Engineering


Dr. Roy Denoon is vice president of CPP Wind Engineering and has been
involved in wind tunnel testing for 30 years, during which time he has
worked on many globally iconic buildings and structures. He is a member
of the ASCE 7 Wind Load Subcommittee, a contributing author to the
ASCE/SEI Prestandard for Performance-Based Wind Design, a co-author of

v
vi AUTHORS

the CTBUH Technical Guide on Wind Tunnel Testing of High-Rise Buildings


and is a CTBUH Fellow.

John Kilpatrick, RWDI


John Kilpatrick is a principal at RWDI. John is currently Wind Engineering
Practice Leader at RWDI, a former chair of the UK Wind Engineering
Society, and is a contributing author to the ASCE/SEI Prestandard for
Performance-Based Wind Design and ASCE 49 Standard for Wind Tunnel
Testing for Buildings and Other Structures.

Sami Matar, LERA Consulting Structural Engineers


Sami S. Matar is an associate partner with LERA Consulting Structural
Engineers (LERA). Since joining LERA in 1995, he has built extensive expe-
rience in a variety of projects and has participated in or led the design of
several of the firm’s high-rise building projects.

Brian McElhatten, Arup


Brian McElhatten is an associate principal in Arup’s Chicago office and
leads the structural engineering group there. He has extensive knowledge
of tall building systems, analysis, and design. During his time at both Arup
and SOM, he has worked on and led the structural design of numerous tall
and supertall buildings around the world. In addition to the Tall Buildings
Committee of SEI, he is actively involved in CTBUH.

Patrick Ragan, WSP


Patrick Ragan is an associate with WSP in Chicago. His design experience
includes towers in the United States, China, and the Middle East, with a
primary focus on the lateral system design and optimization of high-rise
structures. He has authored papers on various subjects including statistical
methods for estimating wind turbine design loads, form-finding analysis of
shell-shaped roof structures, and fluid viscous damper alternatives to con-
ventional steel brace outrigger systems.

CONTRIBUTOR

Alexander W. Jordan, Skidmore, Owings & Merrill


Alex Jordan is an Associate at Skidmore, Owings & Merrill (SOM). He has
experience in the design of tall buildings and airports, and he specializes in
digital design practices.
CONTENTS

PREFACE . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xi
ACKNOWLEDGMENTS. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xiii

1 INTRODUCTION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.1 Purpose . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Scope . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.3 Use of This Manual . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.4 Historic General Design Requirements . . . . . . . . . . . . . . . . . . . . . 3
1.5 Stakeholders . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.6 Nature of Wind . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.7 Limitations. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
2 DESIGN PROCESS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
2.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
2.2 Establish Performance Objectives . . . . . . . . . . . . . . . . . . . . . . . . . 7
2.3 Preliminary Structural Design . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
2.4 Wind Climate Assessment. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
2.5 Wind-Induced Loads and Responses . . . . . . . . . . . . . . . . . . . . . . 9
2.6 Structural Modeling and Analysis . . . . . . . . . . . . . . . . . . . . . . . 10
2.7 Comparison of Results to Acceptance Criteria . . . . . . . . . . . . . . 10
2.8 Wind Optimization Program . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
2.9 Final Design. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
3 PERFORMANCE OBJECTIVES AND
ACCEPTANCE CRITERIA . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
3.1 Introduction. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
3.2 Mean Recurrence Intervals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
3.2.1 Strength: Foundation and Lateral System
(Main Wind Force Resisting System) . . . . . . . . . . . . . . . . . 14
3.2.2 Serviceability: Drift and Displacement . . . . . . . . . . . . . . . . 14
3.2.3 Serviceability: Accelerations and Motion Perception . . . . . 15
vii
viii CONTENTS

3.3 Stability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
3.3.1 P-Delta (Second Order) Effects . . . . . . . . . . . . . . . . . . . . . . 16
3.3.2 Story Stability Coefficient . . . . . . . . . . . . . . . . . . . . . . . . . . 17
3.3.3 Stability Evaluation with P-Delta Analysis . . . . . . . . . . . . 17
3.3.4 Global Stability and Story Stability . . . . . . . . . . . . . . . . . . 18
3.3.5 Stability Acceptance Criteria . . . . . . . . . . . . . . . . . . . . . . . 19
3.4 Strength Evaluation of the Lateral Force–Resisting System . . . . 19
3.5 Building Displacements. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
3.5.1 Overall Building Deflection . . . . . . . . . . . . . . . . . . . . . . . . 20
3.5.2 Story Drift . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
3.5.3 Drift Measurement Index . . . . . . . . . . . . . . . . . . . . . . . . . . 22
3.5.4 Recommended Drift Criteria . . . . . . . . . . . . . . . . . . . . . . . 23
3.6 Nonstructural Elements. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
3.7 Occupant Comfort . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
3.7.1 Acceleration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
3.7.2 Visual and Auditory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
3.8 Project-Specific Performance . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
4 PRELIMINARY STRUCTURAL DESIGN . . . . . . . . . . . . . . . . . . . 29
4.1 Purpose . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
4.2 Preliminary Wind Estimates . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
4.2.1 Along-Wind Response . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
4.2.2 Crosswind Response. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
4.3 Estimation of Building Performance . . . . . . . . . . . . . . . . . . . . . . 30
4.3.1 Preliminary Structural Analysis . . . . . . . . . . . . . . . . . . . . . 31
4.3.2 Strength Checks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
4.3.3 Building Periods and Mode Shapes . . . . . . . . . . . . . . . . . . 32
5 WIND CLIMATE ASSESSMENT . . . . . . . . . . . . . . . . . . . . . . . . . 35
5.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
5.2 Davenport Wind Loading Chain. . . . . . . . . . . . . . . . . . . . . . . . . 35
5.3 Wind Climate: Storm Types and Data Sources. . . . . . . . . . . . . . 36
5.3.1 Windstorm Types . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
5.3.2 Data Sources . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
5.4 Influence of Terrain . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
5.5 Extreme Value Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
5.6 Design Criteria: Mean Recurrence Intervals . . . . . . . . . . . . . . . . 40
6 WIND TUNNEL TESTING . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
6.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
6.2 Triggers for Testing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
6.3 Types of Wind Tunnel Tests . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
6.3.1 High-Frequency Balance. . . . . . . . . . . . . . . . . . . . . . . . . . . 43
6.3.2 High-Frequency Pressure Integration. . . . . . . . . . . . . . . . . 44
6.3.3 Aeroelastic Method. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
CONTENTS ix

6.4 Physical Testing versus Computational Estimates . . . . . . . . . . . 47


6.5 Testing Procedure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
6.5.1 Timeline and Type for Testing . . . . . . . . . . . . . . . . . . . . . . 47
6.5.2 Inclusions and Exclusions. . . . . . . . . . . . . . . . . . . . . . . . . . 48
6.5.3 Required Input Information . . . . . . . . . . . . . . . . . . . . . . . . 48
6.6 Combining Climate and Wind Tunnel Data . . . . . . . . . . . . . . . . 48
6.7 Typical Outputs. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
6.8 Additional Considerations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
6.8.1 Shielding and Influence from Surrounding Buildings . . . . 51
6.8.2 Design Evolution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
6.8.3 Minimum Thresholds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
7 DAMPING . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
7.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
7.2 Inherent Damping . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
7.3 Aerodynamic Damping . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
7.4 Supplemental Damping. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
7.4.1 Direct Damping Systems . . . . . . . . . . . . . . . . . . . . . . . . . . 55
7.4.2 Indirect Damping Systems . . . . . . . . . . . . . . . . . . . . . . . . . 56
7.5 Supplemental Damping: Strength and Serviceability . . . . . . . . . 58
8 STRUCTURAL MODELING AND ANALYSIS . . . . . . . . . . . . . . 59
8.1 Structural Modeling. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
8.1.1 Strength-Level and Serviceability-Level Analysis . . . . . . . 60
8.1.2 Primary Lateral Load–Resisting System and
Nonparticipating Elements . . . . . . . . . . . . . . . . . . . . . . . . . 60
8.1.3 Building Mass . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
8.1.4 P-Delta (Second Order) Effects . . . . . . . . . . . . . . . . . . . . . . 62
8.1.5 Diaphragms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
8.1.6 Foundation Flexibility. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
8.1.7 Panel Zone Deformations . . . . . . . . . . . . . . . . . . . . . . . . . . 64
8.2 Special Considerations for Reinforced Concrete Structures . . . . 65
8.2.1 Expected Strength and Modulus of Elasticity
of Concrete Materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
8.2.2 Stiffness Modifiers and Behavior of Cracked
Reinforced Concrete Structures . . . . . . . . . . . . . . . . . . . . . 66
8.2.3 Simplified Method for Selecting Stiffness Modifiers . . . . . 67
8.2.4 Detailed Method for Selecting Stiffness Modifiers . . . . . . . 67
9 WIND OPTIMIZATION PROGRAM . . . . . . . . . . . . . . . . . . . . . . 71
9.1 Introduction. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
9.2 Building Orientation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
9.3 Building Geometry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
9.4 Holistic Optimization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
x CONTENTS

10 CONCLUDING REMARKS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
10.1 Design Validation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
10.2 Peer Review . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
10.3 Concurrent Research and Future Directions . . . . . . . . . . . . . . . 78
10.3.1 Monitoring . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
10.3.2 Performance-Based Design . . . . . . . . . . . . . . . . . . . . . . . 79
10.3.3 Computational Wind Engineering . . . . . . . . . . . . . . . . . 79
10.3.4 High-Performance and New Materials. . . . . . . . . . . . . . 79
10.4 Closing Remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80

REFERENCES. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81
INDEX . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
PREFACE

The Tall Buildings Committee of the Structural Engineering Institute of


the American Society of Civil Engineers comprises senior leaders active in
the field of tall buildings from both industry-leading professional consult-
ing firms and academic institutions. The goal of this committee is to publish
an industry consensus manual of practice that provides guidelines on the
design and performance of tall buildings for wind effects. This publication
provides recommendations and guidance on tall building design industry
standard practice and approaches to complement current literature, model
codes, and other standards. This publication should be used in conjunction
with local building codes and considered a current consensus document
developed by industry leaders for the design and performance of tall build-
ings for wind effects. However, it is the opinion of the committee that this
consensus will evolve over time and as the industry advances.
Preetam Biswas, P.E., LEED
Director of Structural Engineering
Skidmore, Owings & Merrill
Task Committee Chair—Design and Performance of Tall Buildings for Wind

John Peronto, P.E., S.E., C.Eng, EUR ING, FIStructE, FICE, SECB, LEED AP
Senior Principal
Thornton Tomasetti, Inc.
Technical Activities Division Chair—Tall Buildings

xi
ACKNOWLEDGMENTS

To all the individuals who contributed to the discussion on content:


Maryam Asghari Ramon Gilsanz Viral Patel
Mooneghi Jennifer Goupil Karl Rubenacker
Andrew Bartolini Jeremy Hasselbauer Rob Smith
Audrey Bentz Johnn Judd Seymour Spence
Manotapa Bhaumik Hessam Kazemzadeh John Tessem
Melissa Burton Tracy Kijewski-Correa May Thu Nwe Nwe
Finley Charney Jordan Komp Un Yeong Jeong
Jon Galsworthy Michael Montgomery
Mike Gibbons Deepak Pant

To members of institutes and firms that contributed to the discussion on


content:
Arup Kinetica University of Michigan
ASCE/SEI LERA University of Notre Dame
BCE McNamara Salvia University of Wyoming
CPP MKA Virginia Tech
DCI Rad Urban Walter P Moore
GMS RWDI WSP
Gradient Wind SOM
JHU TT

A special acknowledgement to the members of the Blue Ribbon Panel for


their thorough and detailed review of the contents of this manual:
Abbas Aminmansour, Ph.D.
William F. Baker, P.E., S.E., F.ASCE
Daryl Boggs, Ph.D., P.E.
David Farnsworth, P.E.
William Faschan, P.E., S.E., F.ASCE
xiii
xiv ACKNOWLEDGMENTS

Robert A. Halvorson, P.E., S.E.


Peter Irwin, Ph.D., P.Eng.
Ahsan Kareem, Ph.D., Dist.M.ASCE
Ron Klemencic, P.E., S.E.
Robert Sinn, P.E., S.E.
CHAPTER 1
INTRODUCTION

1.1 PURPOSE

This Manual of Practice is intended to provide best-practice guidelines


for the design of tall buildings for wind. Current building codes and design
standards focus primarily on strength design for ultimate wind loads,
offering little guidance related to the evaluation and establishment of accep-
tance criteria for tall building performance under varying levels of wind
effects. As such, current design practice varies widely across the industry.
The goal of this document is to promote consistency and best practices
within the industry for the design of tall buildings for wind. It is envisioned
to apply to buildings with height greater than 120 m (400 ft) and/or a height
to width aspect ratio greater than 5:1.

1.2 SCOPE

The design recommendations provided in this manual are primarily


focused on the wind design of building structures that are especially tall,
slender, and/or prone to wind-induced movement. A discussion of both
static and dynamic approaches to wind design are presented in subsequent
chapters. Wind design performance objectives are achieved by
• Ensuring structural integrity under ultimate loads;
• Limiting lateral deflections under service loads to prevent permanent
deformations, damage to nonstructural elements, or adverse
effects on the serviceability of the building’s services and vertical
transportation systems;
1
2 DESIGN AND PERFORMANCE OF TALL BUILDINGS FOR WIND

• Preventing excessive wind-induced motions, as well as visual and


acoustical disturbances that may impact occupant comfort;
• Limiting surface wind intensity to improve pedestrian comfort; and
• Achieving these objectives while working to minimize structural
materials and associated embodied carbon and cost.

The manual provides recommendations on achieving the aforemen-


tioned performance objectives, including the selection of the wind loading
mean recurrence interval (MRI) for serviceability, the establishment of
acceptance criteria, guidance on addressing uncertainty in building
structural properties such as stiffness and damping, as well as modeling
of uncertainties in wind climate, wind loading, and other effects.
A concurrent effort to this Manual of Practice is the ASCE/SEI
Prestandard for Performance-Based Wind Design (ASCE 2019). The
Prestandard provides a performance-based approach for the design of
buildings for wind to help resolve the conflicts in performance objectives
that exist in using prescriptive procedures for the wind design and perfor-
mance-based approach for the seismic design of individual buildings.
Among other things, the Prestandard will introduce the use of nonlinear
dynamic analysis for wind design and limited nonlinearity in the main
wind force resisting system elements. Both these topics are outside the
scope of this manual.

1.3 USE OF THIS MANUAL

This manual is compatible with the requirements of ASCE 7-16 (ASCE


2017) for the determination of wind loads and provides additional
guidelines beyond the building code for the establishment and evaluation
of performance objectives when designing tall buildings.
The chapters are ordered in a manner that closely resembles the typical
design approach for a tall building. This usually involves the establishment
of performance objectives and completion of a preliminary structural
design using code-based wind loads, followed by a more detailed assess-
ment of the wind climate and wind-induced response through wind tunnel
testing. The results of the wind tunnel testing program are then used to
refine the preliminary design and complete the final design and construc-
tion documents.
To the extent possible, the chapters are written in a self-contained format
to allow ease of reference regarding specific topics. Significant topics, such
as damping, have dedicated chapters and are cross-referenced from other
chapters to avoid duplicating information.
INTRODUCTION 3

1.4 HISTORIC GENERAL DESIGN REQUIREMENTS

Until the publication of ANSI A58.1-1972 (1972), design for wind loads
was largely governed by local and regional building codes. These codes
typically prescribed wind pressures to be applied to building surfaces.
ANSI A58.1-1972 provided wind load criteria using probabilistically
determined wind speeds and tabulated design load parameters. This prob-
abilistic wind load approach has evolved through revisions made over the
years in ANSI A58.1-1982, and then the various editions of ASCE 7 starting
with ASCE 7-88 (1988), and most recently ASCE 7-16 (2017). The most sig-
nificant changes have been the change to the reference wind speed from
fastest-mile to a 3-s gust in ASCE 7-95, and more recently, the mapped wind
speeds in ASCE 7-10. Until ASCE 7-05, the mapped wind speeds were
based on a 50-year MRI. Importance factors, based on the ratio of 25-year,
50-year, or 100-year velocity pressure to the 50-year velocity pressure, were
used to calculate the wind pressure for the desired MRI from the 50-year
MRI wind pressure based on the building risk category. A load factor of
1.6, obtained through reliability analysis, was used to convert the service
loads to ultimate loads. In ASCE 7-10, wind loads were specified directly
as “ultimate loads,” with a load factor of 1.0. Back-calculations determined
that the equivalent “ultimate” MRIs corresponding to the product of old
importance factors and load factor aligned fairly well with MRIs 300,
700, and 1,700 years. Since ASCE 7-10, ultimate wind speed maps are
published for different risk categories directly representing the 300-year,
700-year, and 1,700-year MRIs (SEAOSC 2010).
Although ANSI A58.1 and ASCE 7 have code mandated the strength
design requirements for wind loads, they have largely remained silent
on serviceability or occupant comfort considerations. Various efforts to
assess the state of the art of wind drift design practices by ASCE and other
organizations such as the American Concrete Institute (ACI) revealed that a
wide range of drift criteria and modeling assumptions are used by design
professionals, and that an industry consensus has not been established. One
of the goals of this manual is to document the industry best practices to
achieve greater consensus and bring increased consistency to the practice.

1.5 STAKEHOLDERS

Whereas the primary target audience for this Manual of Practice are
structural engineers, it is important to recognize the different stakeholders
with a vested interest in the design criteria and to coordinate the building
performance objectives with these stakeholders (Riad 2016). Following is a
listing of significant stakeholders and their potential interests.
4 DESIGN AND PERFORMANCE OF TALL BUILDINGS FOR WIND

• Owners: Owners are the individual(s) or organization that conceives,


develops, and financially supports the project. Owners need to be
involved in major decisions affecting the serviceability of the building
as it may impact the functionality, cost, and value of their property. In
addition, owners may request specific design criteria or design
enhancements beyond what the building code provides.
• Architect: The architect is conventionally the lead design consultant
hired by the owner. The architect works to fulfill the owner’s brief
and coordinates the design among the design consultants. The archi-
tect typically designs or participates in the design of the building
enclosure, vertical transportation system, interior partitions, and
other nonstructural components sensitive to the building’s response
under wind.
• Structural engineer: The structural engineer is engaged by the archi-
tect or owner to design the structure for the building. The structural
engineer’s responsibilities include the design of the building structure
under anticipated loading events including wind loads. The structural
engineer is in the unique position to advise the owner, architect, and
other stakeholders about the anticipated building movements in wind
and resulting motions (such as accelerations) and how to control them
within agreed-on acceptable limits.
• Wind engineering consultant: The wind engineering consultant is
typically engaged by the structural engineer, architect, or owner to
provide the wind loading for use in the design of the building
structure and building enclosure and to determine estimated building
movements, which could impact occupant comfort. The wind
engineering consultant will assess the wind climate and most often
perform wind tunnel tests to obtain the wind loads for the design
of the building structure and to calculate the dynamic response of
the structure.
• Vertical transportation consultant: The vertical transportation consul-
tant designs the building elevator system. Excessive building sway
may impact the operation of elevators. As such, the vertical transpor-
tation consultant has a stake in understanding the building’s drifts
and its dynamic properties to adequately specify the building’s eleva-
tor system.
• Façade consultant: The façade consultant is responsible for the design
and/or specification of the building enclosure. Therefore, accurate
estimation of the wind pressures on the building’s exterior surface
and the wind drifts are essential for the building enclosure design
process.
• Building occupants: The building occupants are not usually part of the
decision-making process during the design of a building unless the
owner occupies the building. However, building occupants are
INTRODUCTION 5

perhaps most impacted by decisions made about the building


performance.
• Neighbors: Neighbors and neighboring properties may be adversely
impacted by a new tall building next to their property. The wind loads
on the neighboring property may be adversely affected as a result of
the new tall building being constructed in the vicinity. Also, the poten-
tial noise caused by wind on exterior elements may become a
nuisance.
• Pedestrians: Pedestrians may be adversely impacted by a new tall
building because wind speeds may increase or be redirected at the
street level in the building surroundings. In some cases, recognizing
the potential impact of a new building on pedestrian comfort and
implementing measures to mitigate high wind speeds will help
facilitate the community’s acceptance of the building.
• General public: The impact of a tall building goes beyond the neigh-
bors or pedestrians, although typically the impact is not wind related.
The impact of a tall building on the general public is often reflected in
the municipal approvals process, which may require additional stud-
ies of the impact of the building on the neighborhood and city.

1.6 NATURE OF WIND

Wind is a dynamic and random phenomenon. Wind speed can be


viewed as a mean value on which random fluctuations or “gusts” (dynamic
effects) are superimposed. The fluctuating component is described in
statistical terms such as standard deviation, power spectral density,
and probability density function. Wind speeds usually increase with
height aboveground, which is a characteristic of a boundary-layer flow
(SP240-2) (ACI 2006).
In addition to this boundary-layer phenomenon, wind has directional-
ity, that is, winds from a particular direction may be stronger or dominate
over wind from other directions. Directionality varies from region to
region, and even from site to site, and is traditionally represented using
a “wind rose” (i.e., a plot showing the variation of wind speed with com-
pass direction).
Further, the concept of a return period (or MRI) is used to calculate the
recurrence interval of winds of a given velocity over time. The return period
is estimated probabilistically using available wind records (or, when avail-
ability is limited, numerical simulation of wind events) and is an essential
concept in the current wind design standards. It is also an important
consideration in the building’s serviceability and performance.
6 DESIGN AND PERFORMANCE OF TALL BUILDINGS FOR WIND

The issue of wind climate assessment is discussed in more detail in


Chapter 5, and broader descriptions of the nature of wind and climate
are available in the literature (Simiu and Yeo 2019, Irwin et al. 2013).

1.7 LIMITATIONS

This manual is intended to provide a framework for the wind design of


tall buildings that is based on the current state of practice in tall building
structural design and wind tunnel testing. The guidelines in this manual
should result in the design of tall buildings that perform adequately in
wind. Because each project may have unique circumstances or require-
ments, engineers and building officials using these guidelines must exercise
their own independent judgment as to the suitability of these recommen-
dations for a specific project. In addition, this manual does not cover
other aspects of tall building design such as, but not limited to, seismic
design, differential shortening, building lean under self-weight because of
differential shortening, and so on. Engineers and building officials must
exercise their own independent, professional judgment as to other design
requirements.
CHAPTER 2
DESIGN PROCESS

2.1 OVERVIEW

This book describes a design process which is intended to be primarily


sequential, beginning with Chapter 3 and ending with Chapter 10 as shown
in Figure 2-1. However, it is understood that most designs do not proceed in
a linear fashion and often require redesign or reanalysis because of changes
by stakeholders or natural design evolution. Furthermore, it will also
frequently be necessary to skip ahead several steps in the design process.
For example, the preliminary structural design will often require the
development of a structural analysis model prior to wind climate assess-
ment or wind tunnel analysis.

2.2 ESTABLISH PERFORMANCE OBJECTIVES

The first step in the design process is the establishment of performance


objectives in accordance with Chapter 3. The performance objectives should
be determined through consultation with the building’s stakeholders and
will serve as targets against which measured performance will be
evaluated.
This book follows conventional tall building design practice in which the
performance objectives fall into two general categories: (1) structural
strength and stability (under rare wind events), and (2) serviceability
(under frequent wind events).
Structural strength and stability requirements are related to life safety
and are typically set by the local building code. Serviceability performance
objectives are related to continued building operation and are not code
7
8 DESIGN AND PERFORMANCE OF TALL BUILDINGS FOR WIND

Figure 2-1. Design process overview.

mandated. As such, it is critical to establish early in the design process the


performance objectives for design considerations such as occupant comfort,
vertical transportation functionality, and damage to nonstructural
elements.
Because the serviceability objectives are not code mandated, for certain
buildings the stakeholders may determine that project-specific perfor-
mance, beyond what is achieved through conventional tall building design
practice, is desired. These guidelines should not be interpreted to prevent
the establishment and use of such project-specific performance objectives.

2.3 PRELIMINARY STRUCTURAL DESIGN

In the preliminary structural design, the engineer must select the


systems, materials, and geometric layout of the structure. Using general
rules of thumb as well as code-based prescriptive methods, the engineer
should determine proportions of overall structural systems and the
individual structural elements. This preliminary design acts as the starting
point for further design refinement and provides a basis for wind tunnel
testing as described in Chapter 6. More detailed information on the devel-
opment of a preliminary structural design can be found in Chapter 4.

2.4 WIND CLIMATE ASSESSMENT

Wind climate analysis is key to any reliable tall building design. The
wind climate analysis is the component of the design process that
determines the probability of a particular wind speed occurring from
any given wind direction. It is also a step that introduces a great degree
of uncertainty owing to the imperfect, and sometimes incomplete, nature
of meteorological data sets.
To increase reliability, different probabilistic approaches may be
required to best fit to the meteorological data for different return periods.
For instance, for short return period events related to serviceability design,
DESIGN PROCESS 9

such as occupant comfort accelerations, a statistical fit that matches


relatively commonly occurring wind events must be used. For strength
design at extreme wind speeds the fit must be of an extreme value type suit-
able for extrapolation well beyond the data record length. For very tall
buildings whose upper portions reach or exceed the height of the atmos-
pheric boundary layer (ABL), it is advised to undertake an upper-level
wind climate assessment to determine wind speeds and directions above
the ABL, because these may have substantial influence on the global
response.
In some jurisdictions, it may be necessary to scale the strength design
wind speeds to local statutory authority requirements. However, in other
jurisdictions, alternative minimum design wind speeds can often be
established using a suitably rigorous statistical analysis. It is recommended
that serviceability assessments be based on a best estimate site-specific
wind climate analysis.
The second part of the wind climate analysis is the transference of wind
data models to the project site, accounting for the upwind roughness for
each direction. Although codes and standards typically provide simplified
methods for assessing upwind terrain, more comprehensive methods may
be used if appropriate. Refer to Chapter 5 for additional information related
to wind climate assessment.

2.5 WIND-INDUCED LOADS AND RESPONSES

The two methods of assessing wind-induced loads and responses are the
use of published data or wind tunnel testing. Wind tunnel testing is
recommended for the design of tall buildings because it is the only current
method that allows the designer to determine the aerodynamic character-
istics for specific shapes and over a full range of wind speeds. Published
data can take the form of information in codes, standards, and published
databases, all of which have been derived from wind tunnel testing.
Some key limitations of published data are that on limited occasions they
reflect the exact geometry of a building under design, the fluctuating
wind-induced response of the structure may not be considered, and they
do not account for the precise surroundings. Wind tunnel testing for a
project overcomes these limitations by providing building and site-specific
information.
Although published data can provide reasonable estimates for the
along-wind response, cross-wind response, which can dominate loads
and accelerations for tall or slender buildings, can only be accurately deter-
mined through wind tunnel tests. This is because wind tunnel tests include
the influence of surrounding buildings, whereas published data may not.
Even relatively small architectural changes can significantly impact a tall
10 DESIGN AND PERFORMANCE OF TALL BUILDINGS FOR WIND

building’s response. Refer to Chapter 6 for additional information on wind


tunnel testing and types of dynamic response.

2.6 STRUCTURAL MODELING AND ANALYSIS

The structural analysis model should represent the properties of the


full-scale building as accurately as is necessary. Appropriate structural
modeling and analysis are important to ensure that tall building behavior
will meet its assumed acceptance criteria once constructed. The model must
include the mass and stiffness of the elements of the lateral system, inclusive
of geometric nonlinearity inherent to the lateral system (P-Delta effects) and
the effects of diaphragms on the distribution of lateral loads. The engineer
must also make assumptions about a number of different building behavior
parameters during the modeling and analysis process. These assumptions
may vary depending on the particular performance objective being studied
and the structural system of the building. Chapter 7 provides more
information about inherent and supplemental damping. Chapter 8 pro-
vides recommendations on best practices for structural modeling and
analysis and the appropriate assumptions that must be made during the
process.

2.7 COMPARISON OF RESULTS TO ACCEPTANCE CRITERIA

The engineer should establish acceptance criteria corresponding to the


performance objectives in accordance with Chapter 3 and compare the
results determined through structural analysis and wind tunnel testing
to the established acceptance criteria to verify acceptable performance.
Where the results do not show acceptable performance, additional design
iterations may be required.

2.8 WIND OPTIMIZATION PROGRAM

Although not required for all designs, there may be a desire to optimize
building response with respect to wind. Chapter 9 provides guidance on the
wind optimization program, which can take the form of either aerodynamic
modification or aerodynamic design. Common optimization strategies
include modification to the building orientation, geometry, and porosity.

2.9 FINAL DESIGN

The final design phase documents the results of the design process. Once
the acceptance criteria are met for the desired performance objectives, the
DESIGN PROCESS 11

final construction documents should be produced. The calculations and


analyses that supported the design process should also be documented
in detail where appropriate. If a peer review is deemed necessary, it should
be completed at this stage of design. An appropriate and experienced peer
reviewer may independently verify the assumptions made and the results
of the design process. Refer to Chapter 10 for more information on the final
design phase and the peer review process.
CHAPTER 3
PERFORMANCE OBJECTIVES AND
ACCEPTANCE CRITERIA

3.1 INTRODUCTION

The focus of this chapter is the establishment of performance objectives


for buildings subjected to wind loading. Where applicable, guidance is pro-
vided on acceptance criteria for use by the structural engineer and other
stakeholders.
The impacts of wind on tall buildings can be divided into two primary
categories: strength and stability, and serviceability.
In general, strength and stability requirements are addressed in the
framework of building codes, ensuring that life-safety requirements
are maintained. Issues related to the strength and stability of tall buildings
are addressed briefly here, including recommendations for global stability
acceptance criteria.
The serviceability performance objectives for tall buildings are primarily
related to controlling lateral movements (both displacements and acceler-
ations) owing to wind. Where the acceptance criteria are not code
mandated, they should be determined by the project’s stakeholders.

3.2 MEAN RECURRENCE INTERVALS

Probabilistic modeling of the wind is critical for determining wind load-


ing on a tall building structure. Sections 3.3 through 3.7 outline various
performance objectives for stability, strength design, drift and deformation
limits, and occupant comfort. Each performance objective should be asso-
ciated with a return period or MRI appropriate for design. The following
sections suggest MRIs for various performance objectives. The MRIs related
13
14 DESIGN AND PERFORMANCE OF TALL BUILDINGS FOR WIND

to serviceability should be considered as suggestions, and as such, they may


vary in actual practice from those listed in this chapter.
In all designs using the ASCE 7 standard or building codes based on the
provisions of ASCE 7, the risk category of a tall building is determined,
which helps set expectations for performance levels for service and strength
design.

3.2.1 Strength: Foundation and Lateral System (Main Wind Force


Resisting System)
Tall buildings must satisfy strength limit states, where structural mem-
bers and components are designed to support the ultimate wind loads and
retain structural integrity. The MRI appropriate for strength considerations
varies based on Building Risk Category. The risk categories and associated
MRIs for strength design per ASCE 7-16 (ASCE 2017) are provided in
Table 3-1.

3.2.2 Serviceability: Drift and Displacement


ASCE 7-16 (ASCE 2017) standard advises that “serviceability limits (e.g.,
maximum static deformations or accelerations) shall be chosen with due
regard to the intended function of the structure.” It goes on to note that
“serviceability shall be checked using appropriate loads for the limit state
being considered.” The code therefore allows significant freedom of choice
with regard to drift and displacement limits and the mean recurrence inter-
val for the assessment of the criteria.
Griffis (1993) recommends a wind load with an MRI of 10 years for
review of wind serviceability limit states of typical buildings. The basis
for this MRI is similar to the average tenancy of office buildings in the
United States, and exceedance of the drift and displacement limits is not
a safety-related issue. Although 10 years may be suitable for interior par-
titions in office buildings, other building components and building types

Table 3-1. Risk Categories with Associated MRIs

Risk Category MRI

I 300 years

II 700 years

III 1,700 years

IV 3,000 years
PERFORMANCE OBJECTIVES AND ACCEPTANCE CRITERIA 15

are expected to last well beyond this time frame, and it is important they can
tolerate movements from larger MRIs.
The commentary to Appendix C of ASCE 7-16 (ASCE 2017), Appendix
CC, provides wind speed maps for MRIs of 10, 25, 50, and 100 years and
notes that a 10-year MRI may be appropriate for a typical building, whereas
50- or 100-year MRIs may be more suitable for drift-sensitive buildings.
For the design of tall buildings, the selection of the appropriate MRI for
drift and displacement is left to engineering judgment and the project
stakeholders. Suggested drift limits for 10-year and 50-year MRIs are pre-
sented in Section 3.5.4. In non-hurricane/non-tropical cyclone regions,
buildings designed to 10-year MRI criteria will in general be able to meet
the corresponding, less stringent limits at longer MRIs. The wind engineer-
ing consultant should provide wind loads and motion predictions appro-
priate for the agreed MRI.

3.2.3 Serviceability: Accelerations and Motion Perception


In North America, a 10-year MRI has traditionally been used for the
determination of occupant comfort. The guidelines were originally based
on early comparisons of wind tunnel predictions and whether complaints
had been received in the occupied buildings. Since then, wind tunnel test-
ing and wind climate analyses have improved, and more recent compar-
isons have been used to refine the acceptable ranges (Irwin and Myslimaj
2008). More recent guidelines have moved to shorter return periods
because these give more of an indication of how often motion is likely
to be perceived in buildings—an important factor in occupant tolerance
of motion. ISO 10137 (2007) is based on a 1-year MRI and has become
the most commonly used guideline internationally. The office guidelines
in ISO 10137 are largely consistent with the previous 5-year MRI guide-
lines published in ISO 6897 (1984), which were based on field experience
of a wide range of building types, including both residential and commer-
cial. The basis for the more onerous residential guidelines in ISO 10137 is
not clear, and these can be difficult to meet for many buildings without
the addition of supplementary damping. This highlights the need for
stakeholder engagement in determining acceleration performance targets,
based on the proposed occupancy and use of the building. For some supertall
and super-slender buildings that may exhibit strong crosswind responses
under moderate wind events, it may be appropriate to examine even shorter
MRIs for comparison with perception thresholds: in many cases, it must be
recognized that to do this would require extrapolation of wind tunnel data
that limits the validity of the comparison. In locations with mixed wind cli-
mates, it is important to compare the influence of different windstorm types
on the likely frequency of motion perception and the magnitude of acceler-
ations during isolated extreme wind events.
16 DESIGN AND PERFORMANCE OF TALL BUILDINGS FOR WIND

3.3 STABILITY

A fundamental performance objective for a tall building under wind


loading is that the lateral system should maintain a minimum margin of
safety against global instability, under all applicable ultimate load combi-
nations. Structural instability is associated with excessive P-Delta (second
order) amplification of deflections and forces. Global and local P-Delta
effects are discussed, and limits for allowable P-Delta amplification are rec-
ommended in the following section.

3.3.1 P-Delta (Second Order) Effects


The lateral stability of a structure is dependent on its lateral stiffness,
lateral movement, and the vertical loads imposed on it. If adequate lateral
stiffness is provided, the structure will maintain equilibrium under hori-
zontal wind-induced movements. Regardless of the amount of lateral stiff-
ness, any vertical compression load acting on a column or wall that has been
laterally displaced will create a geometrically nonlinear action (second-
order effect) as shown in Figure 3-1. These are referred to as P-Delta
(P-Δ) effects and are additional axial, bending, and shear effects that result
from the gravity loads being resisted by a wind-displaced structure. Section
8.1.4 provides recommendations about the destabilizing load P, which
should be considered to capture P-Delta effects under service-level and
strength-level analysis.
P-delta (P-δ) effects refer to the second-order deformation and forces that
occur between a member’s endpoints owing to a destabilizing axial force P.

Figure 3-1. Second-order effects.


PERFORMANCE OBJECTIVES AND ACCEPTANCE CRITERIA 17

For most structures, P-delta amplification is primarily a consideration in


design at the element level. For example, when a perimeter column resists
a horizontal wind load, P-delta effects will amplify the moment from the
horizontal load. However, there are some cases, such as a structural system
that relies on slender moment frame columns to provide stability, in which
the P-delta effect can appreciably reduce the story stiffness. In these struc-
tures, both P-Delta and P-delta effects should be considered in the analysis.
The commentary for Chapter C of AISC 360-16 (AISC 2016) contains two
benchmark problems that may be used to confirm that a given software
program adequately considers these effects. In some cases, it may be nec-
essary to subdivide column elements to capture P-delta effects.

3.3.2 Story Stability Coefficient


In Figure 3-1, the lateral story stiffness, k, is represented by a spring that
has first-order deflection Δ1 under lateral load Vx, with second-order deflec-
tion Δ2 resulting from vertical load P. The destabilizing effect is resisted by
an additional load k*Δ2 in the spring. By statics, PΔ = kΔ2h. Substituting
Δ2 = Δ−Δ1 and k = Vx/Δ1, it follows that the ratio of total deflection to
first-order deflection (the P-Delta amplification) is

Δ=Δ1 ¼ 1=ð1 − PΔ1 =V x hÞ

The quantity PΔ1/Vxh is often referred to as the story stability coefficient,


Q, and appears in various codes; namely, it appears as Q in ACI 318-14. For
any given story, if the story height h, lateral stiffness Vx/Δ1, and gravity load
P are known, the story stability coefficient Q and second-order amplifica-
tion factor Δ/Δ1 may be estimated as follows:

Q ¼ PΔ1 =V x h

Δ=Δ1 ¼ 1=ð1 − QÞ

3.3.3 Stability Evaluation with P-Delta Analysis


Historically, second-order amplification has been evaluated using the
story stability coefficient approach, which lends itself well to hand calcula-
tions. In current practice, such calculations are less common because
modern computer programs can easily account for these effects directly
in a geometrically nonlinear analysis (often referred to as P-Δ analysis when
only small displacements are considered). Still, the concept of the stability
coefficient and the quantification of P-Δ amplification is useful for evaluat-
ing the stability of a structure. When second-order effects are directly con-
sidered in the analysis, a comparison between first-order and second-order
18 DESIGN AND PERFORMANCE OF TALL BUILDINGS FOR WIND

building deflections can be used to evaluate the margin of safety against


instability, and a stability coefficient may be back-calculated as

Q ¼ 1 − Δ1 =Δ

3.3.4 Global Stability and Story Stability


Most tall building structures have substantial cantilever behavior
(movements caused by axial compression and elongation of vertical ele-
ments), and thus the story drift may not be the most effective measurement
of stability. For these structures, linear buckling analysis usually shows
buckled mode shapes with displacement distributed throughout the struc-
ture rather than concentrated at a single story, so it is reasonable to evaluate
stability globally. In this case, Δ/Δ1 may be calculated once in each direction
of interest, based on the roof displacement, rather than at each level.
For structures with lateral systems composed primarily of moment
frames, it may be necessary to evaluate story stability, because the stability
of an individual story could control the stability of the overall structure.
A linear buckling analysis is useful for identifying whether buckling modes
have deformations concentrated in an individual story, or if the deforma-
tions are distributed throughout the height of the structure (Figure 3-2).
If linear buckling analysis is performed, the eigenvalue λ calculated for
each mode represents the factor of safety against instability for that mode

Figure 3-2. Global versus local buckling determined for two structures using
linear buckling analysis: (a) global buckling mode; and (b) story buckling mode.
PERFORMANCE OBJECTIVES AND ACCEPTANCE CRITERIA 19

under the initial gravity load combination considered. The corresponding


stability coefficient is Q = 1/λ.

3.3.5 Stability Acceptance Criteria


Because second-order deflections increase asymptotically as Q
approaches 1.0, building codes have historically set limits well below 1.0
to ensure a sufficient margin of safety against collapse.
The upper limit for Q given in Seismic Design Requirements of ASCE
7-16 [Equation (12.8-17)] is Q ≤ 0.25, and this limit may also be used to dem-
onstrate adequate stability under wind loads, with the gravity loads used in
the calculation of Δ/Δ1 corresponding to the load combinations which
include ultimate-level wind loads.
The limit of Q ≤ 0.25 is equivalent to Δ/Δ1 ≤ 1.33 and implies a factor of
safety of 4.0 against instability. For most high-rise structures, significant lat-
eral stiffness is required to meet other performance objectives, and this limit
will not control the design. For the rare structure where this is not the case, a
higher stability coefficient—and lower corresponding factor of safety against
instability—could potentially be justified. However, the designer should be
aware of the asymptotic nature of instability, and that small changes or
uncertainties in the structure will result in increasingly significant changes
to structural forces and deflections if larger values of Q are allowed.
Stability should be considered in both orthogonal directions correspond-
ing to the primary translational dynamic mode shapes. Torsional stability
should also be considered and may be evaluated by comparing first-order
and second-order displacements under a pure torsional wind case.

3.4 STRENGTH EVALUATION OF THE LATERAL


FORCE–RESISTING SYSTEM

Each structural element of the lateral system should be proportioned to


have adequate strength under all applicable ultimate wind load combina-
tions, including second-order effects. Strength design of structural elements
under combined wind and gravity actions is covered in detail by current
codes including ACI 318-14 and ANSI/AISC 360-16, among others, and
is not discussed further here.
The global stability and associated life-safety performance objectives do
not theoretically require that every element has a demand-to-capacity ratio
less than unity. The building performance could be considered acceptable if
overstressed elements are shown to have adequate ductility so that redistrib-
ution of loads can occur, and the secondary load path is designed for the
redistributed loads. It is the current standard for performance-based seismic
design of buildings to allow controlled inelastic response under ultimate load
conditions (e.g., link/coupling beam yielding, wall tensile yielding, flexural
20 DESIGN AND PERFORMANCE OF TALL BUILDINGS FOR WIND

beam yielding, steel column panel zone deformation, and steel brace yield-
ing). Such behavior under wind loads is beyond the scope of this document
but has been addressed in the recently issued ASCE Prestandard for
Performance-Based Wind Design (ASCE 2019).
One type of material nonlinearity that is considered under current
practice, albeit typically in an indirect way, is concrete cracking. The
selection of these factors will be discussed in detail in Chapter 8. It is
important to note that the second-order forces used for strength design
are dependent on the chosen stiffness–reduction factors.

3.5 BUILDING DISPLACEMENTS

Building displacements under lateral loading can be quantified and


evaluated in several ways. Three commonly used measures, in order of
increasing complexity, are
1. Overall building deflection,
2. Story drift, and
3. Drift measurement index (DMI).

3.5.1 Overall Building Deflection


The simplest measure of describing the displaced shape of a building is
to report the lateral displacement of a single point in plan for each level of
the building. Historically, it has been common to describe the roof displace-
ment under service wind loads as a fraction of the building height, H.
Although the ASCE 7-16 commentary notes that common limits used
for this ratio have ranged between H/600 and H/400, quantitative limits
are not recommended in this Manual of Practice because there are no per-
formance objectives that correlate to this simple measure for tall buildings.
However, it is acknowledged that overall building deflection is commonly
used by structural engineers as a gauge to quickly assess the general behav-
ior of tall buildings prior to evaluating the more rigorous drift measure-
ments described subsequently. Care should be taken when choosing the
point in plan to consider, because results toward the center of the floor plate
may vary substantially from those on the perimeter of the floor plate.

3.5.2 Story Drift


Story drift is the relative horizontal displacement between two adjacent
floors. The story drift ratio is this relative horizontal displacement divided
by the vertical distance between these floors. Figure 3-3 illustrates this
relationship. The points being considered are in the same location in the
horizontal plane (i.e., the points align vertically). As noted in
Section 3.5.1, care should be taken in choosing the point for consideration.
PERFORMANCE OBJECTIVES AND ACCEPTANCE CRITERIA 21

Figure 3-3. Definition of story drift ratio.

Story drifts are important measurements to consider in tall building


behavior, having implications on stability, serviceability, and strength.
Drifts can be composed of deformations caused by flexure, shear, or a com-
bination of both. Depending on the building height and lateral system used
in a tall building, drifts will usually be dominated by either total building
flexure or shear. Figure 3-4 shows a portion of a building undergoing defor-
mation from wind loading, in which Bay 2 is exhibiting primarily flexural
deformation, and Bays 1 and 3 are experiencing primarily shear deforma-
tions. The drift measurement index (DMI) discussed in Section 3.5.3 is a

Figure 3-4. Flexure and shear deformations.


22 DESIGN AND PERFORMANCE OF TALL BUILDINGS FOR WIND

measure that can more appropriately capture the component of drift most
responsible for building damage.

3.5.3 Drift Measurement Index


Although story drift considers only the relative horizontal movement
between adjacent floors, it is helpful to isolate the shear deformation (shear
racking) for a given story by considering the combination of relative hori-
zontal and vertical movements at all four corners of a structural bay or other
area of interest. The DMI was introduced by Charney (1990) and referenced
by Griffis (1993), who noted that it is the racking deformation, rather than
simple story drift ratio, that correlates with damage to cladding and parti-
tions. The DMI may be calculated for each individual bay within a story
using the equations shown in Figure 3-5. As noted in Section 3.5.1, care
should be taken in choosing the location for consideration.

Figure 3-5. DMI for a structural bay.


Source: Adapted from Griffis (1993).
Note: (1) Where there are no differential vertical displacements, the calculation
reverts to the average of the story drift ratios at points A and B. This corresponds
to a lateral system in which the columns do not participate in the lateral system.
(2) Where adjacent columns are linked in the lateral system (e.g., by braces or walls),
the vertical deformation components (D3 and D4) will usually be of opposite sign as
the horizontal deformation components (D1 and D2), and the DMI would be less
than the story drift ratio. (3) Where a participating column and a nonparticipating
column are adjacent to each other, the vertical and horizontal components for that bay
may be additive and the DMI would be greater than the story drift ratio.
PERFORMANCE OBJECTIVES AND ACCEPTANCE CRITERIA 23

Compared to story drift, DMI is a more accurate estimate of the defor-


mations to be experienced by cladding or partitions. Story drift values may
significantly overestimate or underestimate these distortions, depending
on the structural system. If story drift is used as an approximation for
DMI, the accuracy of this approximation should be verified for the building
under consideration (Figure 3-5, Note).

3.5.4 Recommended Drift Criteria


Table 3-2 provides recommended DMI values under service-level winds.
Acceptable drifts for nonstructural elements are a function of material
types, system details, manufacturer requirements, and others. Although
the values shown in Table 3-2 are commonly used in practice, it is important
for the structural engineer to coordinate these with the stakeholders, pref-
erably with input from potential manufacturers during the design phase,
and to make it clear in the contract drawings or specifications what these
assumptions are. Less stringent criteria may be possible but would need to
be evaluated by the appropriate nonstructural component manufacturers
or contractors to understand the cost implications and potential construct-
ability considerations. Information on the cladding panel sizes, partition
detailing, and/or elevator shaft sizes may be needed to appropriately per-
form DMI evaluations for a specific component.
Appropriate MRIs for drift criteria are not code mandated and are at
the discretion of the structural engineer and stakeholders. Larger return
periods like those used for strength design of the main lateral system of
the building may be considered to investigate extreme events and their
effects.

3.6 NONSTRUCTURAL ELEMENTS

Controlling tall building movements under wind load to limit potential


damage to nonstructural elements is an important consideration in the

Table 3-2. Recommended Damage Measurement Index Criteria

Recommended DMI criteria

MRI 10-year 50-year

DMI criteria 1/400 1/300


Note: Story drift ratio may be used instead of DMI in some cases, see discussion in Section 3.5.3.
These criteria are applicable for buildings classified as Risk Category II (10-year MRI is most
commonly used). Buildings classified as Risk Category III or IV may warrant enhanced
performance. See ASCE (2019) for further guidance.
24 DESIGN AND PERFORMANCE OF TALL BUILDINGS FOR WIND

design process. Recommendations are given in Section 3.5.4 and further


discussed in Griffis (1993). The amount of story drift drives the potential
for damage or unacceptable behavior of nonstructural elements. These
drifts are manifest in relative movements perpendicular to the plane of
the elements and/or parallel to the plane of the elements, also referred
to as out-of-plane and in-plane movements, respectively. The in-plane rel-
ative movements can result in racking of the elements and generate shear
deformations.
Some structural lateral systems exhibit movements that contain both
horizontal drifts as well as not-insignificant vertical movements, which fur-
ther exacerbate the racking noted previously. As discussed in Section 3.5.3,
the DMI is usually a more appropriate measure to consider. Closely con-
nected to DMI measurements is the drift damage index (DDI), which is a
measure of how much damage an element has sustained owing to the struc-
tural deformations from wind or seismic actions. Griffis (1993) contains a
thorough discussion on this subject. Because gravity effects can exacerbate
the deflections of horizontal frame members subjected to wind load defor-
mations, it is recommended that the structural engineer consider the total
deflected shape of the frame including both effects when evaluating serv-
iceability issues. Performance objectives and potential damage thresholds
for nonstructural elements should be discussed with and agreed to by
stakeholders during the design process.
The three categories of nonstructural elements usually needing consid-
eration related to tall building drifts are as follows:
• Components and cladding: ASCE 7 defines these as elements of the
building envelope that do not qualify as part of the main wind force
resisting system. In the context of tall building design, this primarily
consists of the façade and cladding system and any secondary struc-
tural supports that attach the cladding to the main structural system.
It also includes roof cladding and any additional parapets or dunnage
elements. Failure of components and cladding elements to perform
properly can result in reduced thermal efficiency, air leaks, excessive
movement or vibration, visual imperfections, falling elements or spall-
ing causing danger below, or major cracks or other failures in the façade
requiring replacement. Cladding materials will play an important role
in drift limits that can be tolerated without special or unusual detailing,
among others. In addition to the considerations at service-level wind
loads, it is also imperative that building envelope design consider
the strength and behavior at large MRIs to prevent failures and/or
determine acceptable levels of damage (Table 3-2).
• Interior partitions: In tall buildings these elements are influenced by
similar concerns and require similar considerations to those of the
components and cladding elements as noted. If building movements
PERFORMANCE OBJECTIVES AND ACCEPTANCE CRITERIA 25

are not properly considered in the design and detailing of the interior
partitions, issues such as visual cracking or audible creaking may be
encountered. Partition materials will play an important role in drift
limits that can be tolerated without special or unusual detailing.
FEMA P-58 (FEMA 2018) and USG Corporation (2014) provide guid-
ance on cracking of interior partitions.
• Vertical transportation: Although shaft alignment in the as-built con-
dition is important for the initial installation of elevator systems,
long-term concerns are usually governed by potential movements
and accelerations under wind loads. Modern elevators have monitor-
ing systems that will force a temporary shutdown in the event of
excessive overall building deflections and associated story drifts or
reduce the travel speed under excessive building accelerations.
Building movements can potentially cause vibration in the elevator
cables that can also lead to temporary shutdowns. Communication
with the elevator consultant regarding anticipated building move-
ment and dynamic properties is important to limit or avoid these
operational issues.

3.7 OCCUPANT COMFORT

Tall buildings, like all structures, sway in the wind. Resonant motion is a
design consideration because of its potential impact on the comfort of the
occupants of the building. However, as human response to motion can vary
significantly from person to person, it is not reasonably possible to satisfy
everyone. When considering the effects of motion on the occupants, the
primary aim is to limit perception of motion under normal conditions
(~0.1 year MRI), to keep occupants comfortable under frequently occurring
events (1-year MRI), and limit discomfort to manageable levels under less
frequent events (10-year MRI).

3.7.1 Acceleration
Although people are sensitive to motions in several ways, one of the
most important ways is the sensitivity to sudden changes in building
motion, which is usually characterized by peak acceleration. (The rate of
change of acceleration, or jerk, is another possible measure but it is not com-
monly evaluated). This peak acceleration occurs as a tall building moves
through a cycle of motion from one extreme to the other, speeding up
through the center of the dynamic swing before slowing down toward
the end of the dynamic swing. At the end of the dynamic swing, as the
motion changes direction, the maximum acceleration occurs.
Building motions caused by wind consist of two components: a static or
sustained action, which is not apparent to occupants but is included in the
26 DESIGN AND PERFORMANCE OF TALL BUILDINGS FOR WIND

estimation of the building drift, and oscillatory or resonant vibration, which


is caused by the dynamic and varying action of wind. This dynamic
response of a tall building to wind excitation is influenced by many factors
as discussed in Chapters 5 and 6.
In general, occupants are more tolerant of discomfort felt infrequently or
for short periods of time, as long as they feel confident about structural
robustness. Early practice focused on accelerations occurring at a 10-year
MRI or a 5-year MRI. ISO 6897 (1984) provided guidelines for a 5-year
MRI based on interpretation of field experience from a large number of
buildings and structures, some that experienced complaints and some that
did not. More recently, a 1-year MRI has been the more common basis for
guidelines [e.g., ISO 10137 and Architectural Institute of Japan (AIJ)
Recommendations (2004)], recognizing the importance of the frequency
of occurrence of perceptible motion in occupant satisfaction. These guide-
lines take into account the building’s natural frequency and its impact on
motion perception. Figure 3-6 shows a comparison of the ISO 10137 and AIJ
criteria. It is observed that ISO residential criteria roughly align with AIJ 90
curves, which indicates that 90% of people will perceive the motion. As
noted, perception of motion should be expected. ISO 10137 guidelines pro-
vide frequency-dependent peak acceleration limits for the 1-year return
period for both residential and commercial occupancy. However, although

Figure 3-6. Comparison of AIJ (2004) and ISO 10137 acceleration criteria
MRI = 1 year.
Source: Courtesy of Arup.
PERFORMANCE OBJECTIVES AND ACCEPTANCE CRITERIA 27

the commercial occupancy guidelines are consistent with ISO 6897, the
stricter residential guidelines are based on factoring the commercial values
by 0.7. This results in a guideline that can be hard to achieve for many build-
ings and more demanding than for many existing successful residential
buildings. The choice of acceleration targets should be made in conjunction
with all stakeholders based on the likelihood, acceptability, and impact of
tenant dissatisfaction or complaints.
Although peak accelerations are normally considered, some people are
prone to discomfort from extended exposure to low-amplitude accelera-
tions (Burton et al. 2015).

3.7.2 Visual and Auditory


Although kinesthetic perception of sway motion in the form of acceler-
ations is the perception mechanism most commonly discussed, there are a
number of other, often more important, cues that may trigger perception of
motion by building occupants. A comprehensive discussion of these can be
found in Burton et al. (2015). The most common of these are visual and audi-
tory cues. Visual cues include swinging lights, moving blinds/curtains,
swaying plants, sloshing liquids, movement of other hung or loosely sus-
pended fittings and objects, and perception of sloping floors. Significant
twisting will often be detected by observers looking outside through win-
dows and perceiving movement relative to surrounding buildings or other
objects. Auditory cues are most commonly nonstructural noise, such as
creaking resulting from the building sway.

3.8 PROJECT-SPECIFIC PERFORMANCE

Occasionally, it may be desirable to provide enhanced performance


beyond the standard performance objectives discussed in the previous sec-
tions. The project-specific performance could take the form of more restric-
tive drift limits, lower acceleration limits, or other criteria. These adjusted
performance objectives may come at a cost premium and should be
discussed with the stakeholders early in the design process.
More restrictive drift limits may be desirable to accommodate brittle fin-
ishes and minimize cracking observed by building occupants in the finished
interior spaces. They may also be required to accommodate smaller joints in
the cladding where desirable. On some occasions, the owner may request
lower acceleration limits for improved occupant comfort.
CHAPTER 4
PRELIMINARY STRUCTURAL DESIGN

4.1 PURPOSE

In all but the simplest cases, the structural design process is iterative
rather than linear. This is particularly true for the design of tall buildings
for wind, in which accurate wind loads are necessary to proportion the
building’s lateral system. Wind loads, however, cannot be accurately
estimated before the lateral system has been proportioned and the associ-
ated dynamic properties have been calculated. This chapter addresses the
initial steps in this iterative process that are necessary before wind tunnel
testing can take place.

4.2 PRELIMINARY WIND ESTIMATES

4.2.1 Along-Wind Response


A first step in the preliminary structural design of a tall building is
typically to estimate the design wind loads in the along-wind direction
(parallel to the wind) using either code provisions such as those in
ASCE 7-16 (ASCE 2017), or engineering judgment from past wind tunnel
testing.
For many tall buildings, code-estimated wind loads in the along-wind
direction may be conservative because shielding from other buildings
and wind climate directionality are not directly considered. A site-specific
wind climate analysis may be used to improve estimates of wind-induced
structural loads and responses. The ASCE 7-16 (ASCE 2017) commentary
contains a simplified method for interpolating between upwind
29
30 DESIGN AND PERFORMANCE OF TALL BUILDINGS FOR WIND

roughnesses, an approach based on the work of Deaves and Harris (1978).


This approach can be used along with, or without, a directional wind
climate analysis (described in Chapter 5) to more accurately assess design
wind speeds at a project site.

4.2.2 Crosswind Response


For many tall buildings, response in the crosswind direction
(perpendicular to the wind) will exceed response in the along-wind direc-
tion. Crosswind vibration in buildings is most commonly caused by vortex
shedding. This is a phenomenon in which vortices are created and shed
alternately from each side of the building. The rate at which these vortices
are shed is a function of the wind speed, building width, and building
shape. Where the frequency of the vortex shedding is close to the natural
frequency of the building, the structure will experience crosswind reso-
nance leading to, depending on the building and climate, large accelera-
tions and/or loads. For very tall, slender buildings with low natural
frequencies these large responses can occur at relatively low MRIs, with
implications for occupant comfort. Even for stockier buildings, the peak
crosswind response can occur at wind speeds below that associated with
the strength design MRI wind speed and should be accounted for in the
design.
Methods of preliminary estimation of crosswind response are found in
some codes and standards such as AS/NZS 1170.2:2011 (2011) or NRC
(2015). These are, however, relatively crude estimates for simple building
shapes without significant influence from surrounding buildings. Other
publicly available sources of aerodynamic data for the preliminary design
of tall buildings include the NatHaz Aerodynamic Load Database (NALD)
maintained by Notre Dame University. This resource provides estimates
for base reactions and accelerations for various building shapes under
along-wind, crosswind, and torsional responses, based on a database of
wind tunnel results for isolated buildings. The NALD and other data-
base-enabled design (DED) tools related to wind loads are available at
https://www.vortex-winds.org/.
Other sources of wind-loading data can be used to refine these estimates,
such as wind tunnel test data for buildings of similar shape and/or
surroundings. Often this information is not available in the public domain,
but wind engineering consultants may hold this data in-house for use in this
type of early analytical prediction.

4.3 ESTIMATION OF BUILDING PERFORMANCE

The next step for the designer is to lay out and proportion a structural
system to meet key performance objectives based on the estimated wind
PRELIMINARY STRUCTURAL DESIGN 31

loads. A preliminary structural analysis should be performed to estimate a


building’s lateral stiffness with reasonable accuracy and to ensure that the
lateral stiffness is sufficient to meet the drift performance objective under
the preliminary estimates for service-level wind loads.
It is equally important to estimate a building’s mass realistically, because
mass and stiffness (and their spatial distribution) are the fundamental
properties affecting a building’s periods of vibration and the associated
mode shapes. The periods, mode shapes, and assumed damping ratios
are combined with wind tunnel test results by the wind consultant to deter-
mine wind loads and building response. These attributes, along with wind
climate, the site’s surroundings, and the shape of the building, compose the
key variables that determine the structural loads to be used in the final
design.

4.3.1 Preliminary Structural Analysis


For preliminary analysis, the primary goal is to generate sufficiently
accurate structural properties for wind tunnel testing. Recommendations
for this level of analysis are listed in this section. A more detailed discussion
of structural modeling and analysis follows in Chapter 8.
The structural analysis model may be relatively simple at the preliminary
stage, but at a minimum should include the primary lateral system elements;
these elements should be proportioned so that the maximum allowable
movements from Chapter 3 are not exceeded under serviceability conditions.
Gravity elements (those that are not part of the lateral system) may be
omitted for the analysis model, but gravity loads, including the self-weight
of any structural elements not included in the model, need to be included
to allow
• Estimation of building mass for calculation of building periods and
mode shapes,
• Evaluation of P-Delta effects, and
• Evaluation of resistance to overturning.

P-Delta effects should always be considered, and a comparison of the


deflection including P-Delta effects, Δ, to the first-order deflection, Δ1,
should be made to evaluate the global stability performance objective from
Section 3.3.5.
For concrete construction, appropriate stiffness modifiers should be
used to account for cracking. The simplified approach of selecting stiffness
modifiers from a table (see Section 8.2.3) is typical for preliminary analysis.
The extent of concrete shear wall tensile cracking is particularly
important and should be closely scrutinized. If extensive wall tensile
cracking is observed under serviceability conditions, building drifts may
increase significantly, and revisions to the structural system will likely
32 DESIGN AND PERFORMANCE OF TALL BUILDINGS FOR WIND

be required to meet the drift performance objective. For most tall concrete
shear wall buildings, wall tensile cracking under service loads should be
limited to small localized areas or avoided completely.
Various other topics affecting lateral stiffness in an analysis model are
discussed in Chapter 8, including panel zone deformation and foundation
flexibility, and these issues may or may not warrant consideration at the
preliminary stage.

4.3.2 Strength Checks


It is not necessary to perform detailed member design (e.g., selecting
concrete reinforcement to meet strength checks) at this stage. The focus
should be to determine whether key elements of the lateral system, such
as link/coupling beams, outriggers, shear walls, and so forth, are within
code-mandated design limits. If they are not, these elements should be
reproportioned or the structural system should be revised, or both.
The feasibility of the foundation system should also be considered.
If foundation uplifts are present, the ability of the foundation system to
resist these uplifts should be confirmed.

4.3.3 Building Periods and Mode Shapes


The preliminary structural analysis may be considered complete when
the following are confirmed:
• Structural mass has been accurately estimated;
• The lateral system has been selected, and preliminary sizes have been
assigned for lateral system components;
• Applicable structural stiffness reductions, especially concrete
cracking, have been considered;
• P-Delta effects have been considered;
• Structural drifts have been verified not to exceed the target drift limits
defined for the performance objective under estimated serviceability
wind loads;
• Key lateral system elements can be designed; and
• Foundation elements can be designed.

At this point the structural mass and stiffness in the analysis model may
be deemed sufficiently accurate to be used to determine the building peri-
ods and mode shapes to be reported to the wind tunnel. Ultimately, as the
design progresses past the preliminary stage and wind estimates are
replaced with wind tunnel loads, it is almost certain that the building
periods and mode shapes will evolve. This is especially true if a wind opti-
mization program (as described in Chapter 9) is pursued to reduce wind
effects. In either case, the revised structural properties should be discussed
with the wind engineering consultant to determine if it is necessary to
PRELIMINARY STRUCTURAL DESIGN 33

update the wind loads to be used for design, which could lead to further
changes in the structural properties or the requirement of a new wind
tunnel test. At the preliminary stage, however, the designer’s goal is simply
to estimate the structural properties with a level of accuracy reflective of the
information available, which should minimize the number of future design
iterations.
CHAPTER 5
WIND CLIMATE ASSESSMENT

5.1 OVERVIEW

An understanding of the strength and directionality of the wind climate


local to the project site is crucial to ensure the safe and efficient design of a
tall building. This chapter outlines the role that wind climate plays in
determining wind loads on a tall building and discusses best practices
for obtaining and processing data that describes the wind climate, the
statistical modeling of the data, and how it is combined with the results
of wind tunnel testing.

5.2 DAVENPORT WIND LOADING CHAIN

A convenient approach for evaluating wind loads and the ensuing


wind-induced responses of tall buildings can be described by way of the
Davenport wind loading chain. This approach was developed by Alan
G. Davenport over a lifetime of work in the field of wind engineering
(Davenport 1977, 1982), and it ties together the various concepts used to
evaluate wind actions on structures using a rational “chain of thought”
process. The wind loading chain is shown in Figure 5-1.
The basis of the analogy to a chain is that predictions of wind loads and
effects are only as sound as the strength of the links in the chain. It is there-
fore rational to ensure that each link of the chain (and the understanding of
each component’s impact on the predicted response) is equally strong. The
first link in the chain, and the link contributing the largest amount to the
variability of predictions of wind-induced response, is wind climate.
35
36 DESIGN AND PERFORMANCE OF TALL BUILDINGS FOR WIND

Figure 5-1. Davenport wind loading chain.

The following sections review windstorms, the influence of terrain on


local wind speeds, statistical approaches to integrate wind climate with
wind tunnel measurements of response, and the recurrence intervals for
various limit states.

5.3 WIND CLIMATE: STORM TYPES AND DATA SOURCES

The type of windstorm contributing to the wind hazard at a site will vary
with the geographic location. This is important to the design of tall build-
ings for two primary reasons: (1) different windstorms possess different
mean and gust wind profiles; and (2) it has been shown that statistical
analysis of wind speed and direction data, especially through extreme
value analysis, should be completed on data sets separated by storm type.

5.3.1 Windstorm Types


Windstorms can be classified into the following types:
• Synoptic low-pressure systems: These are large-scale storms, with
their stronger winds usually thought of in terms of gale events.
• Tropical cyclones: Depending on severity and location, these may be
referred to as tropical depressions, tropical storms, hurricanes,
typhoons, or cyclones. They are associated with extremely high wind
speeds, heavy rainfall, and coastal storm surges and swells.
• Thunderstorms: These are much more localized storms in which the
highest wind speeds result from downdrafts impinging on the
ground and spreading out. Consequently, their highest wind speeds
are often close to the ground, and they have very different profiles and
temporal characteristics to synoptic-scale storms.
• Tornadoes: These are small-scale, fast moving windstorms that can
produce extremely high wind speeds with very high rotational flows.
• Topographic or thermally driven winds: These are winds created by
high density air flowing from a high elevation down a slope under the
force of gravity. Special regions for these types of winds are identified
in the ASCE 7 wind maps.
WIND CLIMATE ASSESSMENT 37

A discussion of the local wind climate should be included as part of the


submission from a wind tunnel laboratory. This discussion should include
a description of the prevalent windstorm type(s) that impact the project site,
how these storms have been reflected in the subsequent wind tunnel
testing, and how they have been treated in the statistical analysis of the
wind climate. The treatment of the wind climate may differ for strength
and serviceability design requirements. The wind tunnel consultant should
clearly document that the local wind climate characteristics and classifica-
tion have been incorporated into the wind loads and effects predicted for
the building.

5.3.2 Data Sources


Data that describe the local wind climate can be either measured or
simulated. Measured data are typically obtained from surface weather
stations, where the data have often been archived for many decades. The
wind velocities measured at surface height, typically 10 m (33 ft) above
ground level, are extrapolated to building height or some other clearly
defined reference height using models of boundary layer winds. Wind
climate analyses conducted for tall buildings typically use measured data
as their primary data set, although it may be augmented by other data
sources. Surface data are commonly recorded at 1 h intervals, although
some stations record data once every 3 h. Many international stations
may have sparser records, depending on the location. In addition, many
weather stations may have a long history of seemingly consistent data,
yet the anemometer may have been relocated or replaced with one of a dif-
ferent type, or vegetation or architectural build-ups may have occurred,
resulting in a change in its effective exposure, height above ground, and
its gust response characteristics. Such effects must be identified, and the
appropriate analysis must be executed to convert all readings to a consis-
tent 3 s gust speed at a height of 10 m (33 ft) in open country. Care and atten-
tion must be given to stations that do not have continuous records, and
corrections should be applied as appropriate.
Simulated wind data is often used in regions having frequent hurricane
or cyclonic wind events to overcome the limitations of measured data.
Reliable measurements of hurricane winds are also hampered somewhat
by the robustness of instrumentation in extreme events. Monte Carlo meth-
ods are used to simulate a large number of time histories of hurricane events
(typically between 10,000 and 100,000 events) to increase the population of
events used to predict extreme winds, which also enhances the reliability of
the statistical analyses for ultimate limit state design.
38 DESIGN AND PERFORMANCE OF TALL BUILDINGS FOR WIND

5.4 INFLUENCE OF TERRAIN

In tall building design practice, the structural engineer works with the
client to identify specific performance objectives for serviceability and
strength limit states. As noted in Chapter 2, a key design parameter is wind
speed—pressure varies as a function of wind speed raised to the power of 2,
but building response may vary as a function of speed to higher powers,
typically in the range of 2.3 to 4. Site wind speeds for design are affected
by both near-field and far-field effects, related to the proximity of surround-
ing buildings and the aerodynamic roughness of the terrain, respectively.
Near-field effects are understood but difficult to quantify accurately
using an analytical approach. It is understood that neighboring buildings
having massing similar to the target building can provide shelter if located
in close proximity, but interference effects such as turbulent wakes or
channeling by multiple buildings can also be more onerous for design.
Interference is not typically addressed in codified approaches to calculating
wind-induced loads because of the dependency of the wake effect on
spacing and upwind building massing and geometry. Because near-field
effects are best understood through wind tunnel modeling, they form
the focus of Chapter 6.
Far-field effects related to the aerodynamic roughness of the ground
influence both the mean and gust wind velocity profiles at a site. In nature,
the progression from one terrain type to another is usually gradual, and the
range of aerodynamic roughness length covers several orders of magni-
tude. For tall buildings it is recommended that the influence of terrain
on velocity and turbulence be evaluated using rational approaches
complementary to code, such as the model of the planetary boundary layer
provided by the Engineering Sciences Data Units (ESDU) based on work by
Harris and Deaves (1980).
Changes in local topography can affect wind speeds in the near and far
field. Slopes with grades greater than about 5 degrees can locally increase
wind speeds near the crest of hills and escarpments, particularly close to the
earth’s surface. Topographic effects are calculable (e.g., the Kzt factor in
ASCE 7), and analytical methods are provided in all codes of practice.
In complex terrain, topographic effects are better quantified through the
use of reduced scale wind tunnel models (1:2000 to 1:5000 scale) or compu-
tational wind engineering.
The wind consultant’s report should include an explanation of the
techniques used to assess terrain influence and provide a description of
the wind tunnel profiles used in the simulations of wind effects.
WIND CLIMATE ASSESSMENT 39

5.5 EXTREME VALUE ANALYSIS

The determination of an applicable design wind speed or pressure is


conventionally based on extreme value analysis techniques. This analysis
involves fitting a statistical distribution to independent maxima. Most
maxima extraction techniques fall into the following categories:
• Annual maxima: The highest magnitude wind speed in a calendar
year (nominally January 1 to December 31) is extracted for each year
in the available period of record. This limits the number of maxima for
fitting to the number of years in the period of record.
• Monthly maxima: The highest magnitude wind speed in each month
of each year is extracted. For example, if there are 30 years of data in
the historical record, then 12 × 30 = 360 maxima are available for
fitting.
• Method of independent storms: In this method, independent maxima
are extracted using a two-pass algorithm that first identifies indepen-
dent events and then identifies the maxima within the events. This
significantly increases the number of independent maxima to upward
of 80 or 90 per year.

The method of independent storms is considered more robust than the


annual or monthly maxima methods because statistical independence is
implicit in this approach. The wind consultant’s report should include a
description of the extreme value analysis techniques used as well as a
discussion regarding the maxima extraction techniques.
Before fitting, extracted maxima should be classified according to storm
type (e.g., synoptic, thunderstorm, and so forth). It is well established that
in mixed climates, where multiple storm types contribute to the extreme
wind climate, individual extreme value fits should be conducted for each
storm type. For example, a site may be subject to thunderstorm, synoptic,
and tropical cyclone winds. Extreme value fits should be conducted on the
extracted maxima individually and then recombined to determine the over-
all risk.
For serviceability limit states, it is common practice that the wind
demand determined for any given MRI will be based on the wind consul-
tant’s statistical evaluation of the extreme wind climate, which is typically
at or below the code-prescribed wind speeds or pressures. This analysis also
allows for directional wind effects to be included. For strength design, wind
loads are typically provided at the MRI of interest based on the code-
prescribed wind speed or pressure. In areas where no code-prescribed wind
speed or pressure is available, or where the wind consultant and design
team feel that the code is inappropriate for the site or building under study,
a detailed wind climate report can be prepared by the wind consultant and
40 DESIGN AND PERFORMANCE OF TALL BUILDINGS FOR WIND

presented to the local building code official for approval. Sections 26.5.3
and C26.5.3 of ASCE 7-16 (2017) describe the procedures that should be
followed for estimating basic wind speeds from regional climatic data.
A peer review of the analysis may be recommended.

5.6 DESIGN CRITERIA: MEAN RECURRENCE INTERVALS

Probabilistic modeling of the wind hazard is critical for determining


the demand on a tall building structure. In most designs, the occupancy
category of a tall building is determined, which sets expectation on
performance levels for service and strength design.
Various performance objectives for stability, occupant comfort, drift
limits, and strength design are outlined in Chapter 3. Each performance
objective is associated with an MRI appropriate for design considerations.
CHAPTER 6
WIND TUNNEL TESTING

6.1 OVERVIEW

This chapter outlines the role that wind tunnel testing plays in
determining wind loads on a tall building and accounting for the uncertain-
ties that are contained in code-based wind analysis. The following sections
discuss triggers for testing in the wind tunnel, the types and benefits of each
test, how to identify the strengths and weaknesses of physical and
computational testing, and test inputs and outputs. The benefits of wind
tunnel testing for the design of a tall building are highlighted.

6.2 TRIGGERS FOR TESTING

Satisfying the ultimate strength demand for a tall building is not a


guarantee that serviceability limit states (commonly, building motion
and drift) will be met. For example, even in high seismic zones where
strength design is governed by seismic requirements, this is no guarantee
that occupant comfort under wind excitation will be achieved. This results
from the random nature of the applied wind force having energy spread
across a broad range of natural frequencies and the sensitivity of a build-
ing’s response at varying frequencies. The sensitivity of a building’s
response is a function of its dynamic structural properties, its geometric
form, the velocity and turbulence of the approaching wind (which includes
the effects of local topography), and interference effects from nearby
structures.
Any combination of the aforementioned variables may serve as the
triggers for wind tunnel testing. However, if more than one of the following
41
42 DESIGN AND PERFORMANCE OF TALL BUILDINGS FOR WIND

conditions are fulfilled, the design team should seek specialist advice from a
wind consultant and consider the use of a wind tunnel from the early stages
of design. Items that might affect the wind effects on a tall building include
the following:
1. Height and slenderness ratio: Crosswind response, which is not
covered by many codes, will often govern the wind response of build-
ings with height greater than 120 m (400 ft) and/or a height-to-width
aspect ratio greater than 4:1.
2. Irregular plan forms not covered by code provisions: Examples
include buildings with unusual architectural features, such as
tapered, twisted, or offset stories; rooftop crown features; and build-
ing appurtenances.
3. Buildings with plan forms known to be susceptible to crosswind
loading or vortex shedding should be evaluated using the wind
tunnel approach.
4. The site location: When a tall building is located on a site where buf-
feting in the wake of upwind obstructions or channeling effects
caused by proximity to neighboring buildings is anticipated, special
consideration is warranted.
5. Linked tower structures: Where multiple towers have linking struc-
tures, such as skybridges, it is important to understand the relative
movements between the towers and/or the load transfer through
the linking elements.
6. Unusual structural dynamic properties or asymmetric structural
systems that may not be well covered by code provisions.

6.3 TYPES OF WIND TUNNEL TESTS

For the design of tall buildings, stakeholders typically seek distributed


wind loads for the design of the main wind force resisting systems
(MWFRS) and foundations, and predictions of building drifts and motions
for serviceability and occupant comfort. To evaluate the wind-induced
responses of a tall building in the wind tunnel, there are two basic types
of wind tunnel tests: (1) static (rigid) model tests, and (2) elastic model tests.
In general, static model tests are undertaken using one of two methods:
the high-frequency balance (HFB) approach, in which wind forces
(moments and shears) are measured directly at the model’s base, or by
using integration of simultaneously measured surface pressures known
as high-frequency pressure integration (HFPI). The model is rigid for both
approaches. The elastic model technique, known as the aeroelastic model
method, incorporates a scaled elastic model of the tall building. Brief
explanations of the wind tunnel test types are provided subsequently.
WIND TUNNEL TESTING 43

6.3.1 High-Frequency Balance


The HFB is a device capable of measuring up to six components of force
and/or moment. The balance is typically integrated with a lightweight
wind tunnel model to develop high stiffness, yielding natural frequencies
of vibration of the combined model/balance system that are beyond the
range of interest for post-test analyses, hence the term, high-frequency
balance (Figure 6-1). Mean and fluctuating loads integrated over the rigid
model’s surface are measured directly by the balance. The measured,
applied base moments are combined with the estimated structural modal
properties to determine design wind loads. Care must be taken by the wind
consultant to ensure that corrections are applied to the measured data
where the building mode shapes deviate from a linear variation with
height. This is an aerodynamic type model test, in which the effects of
the dynamic properties are incorporated analytically.
The high-frequency balance approach is well suited to wind tunnel
investigations early in the design process because the models are relatively
simple to build and instrument. This method is the most suitable for form-
finding workshops in which geometric form is being evaluated and early
design feedback is required for evaluating different structural systems.

Figure 6-1. HFB model: One Bangkok Tower.


Source: Courtesy of RWDI.
44 DESIGN AND PERFORMANCE OF TALL BUILDINGS FOR WIND

6.3.2 High-Frequency Pressure Integration


The HFPI technique is another aerodynamic model test that uses a rigid
pressure-tapped model, typically the same model used for the determina-
tion of local cladding pressures (Figure 6-2). For this reason this approach is
best suited for a building whose geometry and external features that may
impact building aerodynamics are relatively finalized. Rather than measure
forces directly, the simultaneous pressure integration approach instantane-
ously combines pressures over the envelope to derive mean and fluctuating
components of the applied wind pressures. The model must contain a
density of pressure taps on the surface of the model sufficient to capture
key aerodynamic flow characteristics.
The resonant component of the wind-induced response is calculated by
directly integrating local surface pressures with the dynamic modal proper-
ties. This also permits higher modes of vibration to be included in the
dynamic analysis, something that is not possible with a typical HFB test.
This may be important for tall buildings with extremely long fundamental
periods of vibration (i.e., greater than 10–12 s), in which secondary sway
modes may be sensitive to wind-induced excitation.

Figure 6-2. High-frequency pressure integration model: Torre Shyris 18, Quito,
Ecuador.
Source: Courtesy of CPP.
WIND TUNNEL TESTING 45

The pressure integration technique allows the wind force distribution


with height to be resolved with improved precision compared to the
high-frequency balance approach. This is beneficial in predicting the distri-
bution of quasi-static (background) loads among floors and resonant
response when torsion is important. The pressure integration technique
is also useful for tall buildings with different structural systems over its
height, for example, a concrete tower with a lightweight steel rooftop crown
feature. The integration of pressures can be undertaken to inform the design
of smaller architectural/structural features such as rooftop crowns and
winter gardens, among others, and does not require a separate test. Where
higher modes of vibration may be excited by the wind, it is recommended to
verify wind-induced response predictions from early HFB studies using the
simultaneous pressure integration approach where possible.
For particularly tall and slender buildings, or those with aerodynami-
cally significant architectural elements that cannot be pressure tapped,
the high-frequency pressure integration approach may not be possible
because of the large number of pressure taps required and the inability
to extract the tubes associated with these within the small plan form of
the building. Similarly, for buildings with very complex architecture, it
may not be possible to include a sufficient number of pressure taps to
adequately define the instantaneous pressure fields over the building
envelope.

6.3.3 Aeroelastic Method


An aeroelastic model is designed to have scaled geometry, mass, stiff-
ness, modal displacements, and damping characteristics that represent
the full-scale prototype of the building (Figure 6-3). In other words, the
model will respond to applied wind forces elastically by vibrating in the
same fashion as the actual building. The benefit of the elastic response is
that aerodynamic forces that may occur owing to large-amplitude motion
of the building will be captured in the measurements of the model’s
response. Aeroelastic models may be simplified to evaluate only the funda-
mental sway modes, or they may be more complex to permit measurement
of the response of the building in higher modes or in torsion.
An important aeroelastic response occurs when the motion of a structure
results in a change in the wind loading. If the resulting change in wind
loading acts in phase with the structure’s velocity, it may reduce the
aerodynamic stability of the structure. It is convenient to include these
velocity-induced forces in the response equation of the structure, and the
commonly used term for these forces is aerodynamic damping. For most tall
buildings, aerodynamic damping will be positive. This can be beneficial in
cases in which the predicted wind-induced response(s) marginally exceeds
the performance criteria established by the design team, and the positive
46 DESIGN AND PERFORMANCE OF TALL BUILDINGS FOR WIND

Figure 6-3. Full aeroelastic model: 432 Park Avenue.


Source: Courtesy of RWDI.

aerodynamic damping will lead to small reductions in the response. For


particularly slender buildings, aerodynamic damping in the crosswind
direction may be negative, particularly around critical wind speeds for
vortex shedding. This effectively reduces the total damping of the building.
This is not an effect that may be measured using rigid model methods
approaches, although the results from HFB and HFPI testing can give an
indication that the building is operating in a wind speed range where this
is possible. Where negative aerodynamic damping may occur, an aeroelas-
tic model test is recommended. When designing for ultimate limit states,
the controlling response may occur at shorter MRI wind speeds. This would
be as a result of vortex shedding, and the impact of negative aerodynamic
damping needs to be considered.
The aeroelastic study is the most cost- and time-intensive wind tunnel
test for a tall building. If deemed necessary, it is often performed in the latter
stages of design. The aeroelastic model can provide a final high-fidelity con-
firmation of the wind-induced response when required, although care
needs to be taken with buildings with complex coupled modes to ensure
that the necessary parameters are adequately modeled (ASCE 49-12).
WIND TUNNEL TESTING 47

6.4 PHYSICAL TESTING VERSUS COMPUTATIONAL ESTIMATES

Wind tunnel testing is the established tool used to model wind flow
around buildings and structures to develop design data for cladding and
structural wind loads. It is also used to conduct pedestrian-level wind
studies and assess pollutant dispersion. Wind tunnel modeling provides
a good representation of both mean and turbulent wind effects, and there
are a number of full-scale validation studies of model-scale wind tunnel
predictions (Li et al. 1998, Kijewski-Correa et al. 2006) which substantiate
the approach.
Computational wind engineering (CWE), typically using computational
fluid dynamics (CFD) techniques, can be a useful tool for simulating flow
behavior and some flow physics that cannot be achieved in a wind tunnel.
However, CWE modeling requires specially trained analysis, and tech-
niques have not yet reached the stage of development at which they can
routinely and accurately predict separation and turbulent eddies and gusts
within the urban environment, as well as flow around buildings. Although
significant advances are being made in the field of CWE, at present there are
limited CWE studies that provide accurate validation of global wind loads
for structural design for simple isolated buildings or clusters of buildings.
Given the massive computing resources that would be required to
adequately model atmospheric flows through a typical urban environment
to predict loads on an architecturally complex building, it will likely be a
significant number of years before CWE is capable of predicting overall
wind loads on buildings with the necessary accuracy. In the meantime,
wind tunnel testing is the recognized approach to determine the various
wind effects on buildings.

6.5 TESTING PROCEDURE

A number of standards and guidelines for wind tunnel testing proce-


dures are available, such as ASCE 49-12 (ASCE 2012), which is referenced
by ASCE 7-16 (ASCE 2017), the ASCE Manual of Practice 67 (ASCE 1999),
and the Council on Tall Buildings and Urban Habitat (CTBUH) guideline,
Wind Tunnel Testing of High-Rise Buildings (Irwin et al. 2013). The following
section reviews wind tunnel model types and what stages of design they are
appropriate for.

6.5.1 Timeline and Type for Testing


The design of a tall building typically progresses through three distinct
design phases: concept and schematic design, design development, and the
technical design or construction drawings phase. During the concept and
schematic design phase, high-frequency balance tests are most often used
48 DESIGN AND PERFORMANCE OF TALL BUILDINGS FOR WIND

to supplement the building code for the initial assessment of overall wind
loads and responses. The HFB approach permits the most economical and
quickest investigation of wind effects and provides rapid feedback to the
design team. Once the architectural design has been finalized, the overall
wind loads and responses can be determined using either the HFB or
HFPI method, depending on the complexity or slenderness of the architec-
ture. Aeroelastic tests are normally specified after HFB or HFPI testing if the
results from these tests indicate that aerodynamic damping effects may be
significant for design.

6.5.2 Inclusions and Exclusions


Existing buildings and buildings currently under construction should be
included in the wind tunnel models. Where future development is known,
it should be considered for inclusion as an additional test configuration if
the wind tunnel consultant believes it may significantly affect the results.
Some local planning authorities maintain a list of developments that
have received planning consent. This information can be used to guide
the design team regarding future developments to include in the additional
tests. If this information is not available, the design team in conjunction
with the wind consultant should reach a consensus regarding future
development to include in the additional tests.
As noted in Section 6.8.3, specific testing of alternative surroundings
may be required to address minimum thresholds required by code.

6.5.3 Required Input Information


To undertake wind tunnel studies for structural loads, several inputs are
required by the wind tunnel consultancy from the design team.
Wind tunnel models may be constructed using two-dimensional (2D)
plans, sections, and elevation drawings, and are commonly built of
balsa wood, high-density foam, and thermoplastic resins. However, it is
now more common for the wind engineering consultant to be provided
information in three-dimensional (3D) format, and 3D rapid prototyping
technologies are increasingly used for model manufacture.
Analysis of wind tunnel data requires the following structural informa-
tion as basic inputs: mass distributions, stiffness centers and mass moments
of inertia, modal displacements provided about a reference coordinate,
frequencies of vibration, structural damping ratios, drawings of the struc-
tural model coordinate origin, and orientation of the structural axis system.

6.6 COMBINING CLIMATE AND WIND TUNNEL DATA

Several approaches may be used to incorporate wind climate statistics


into the prediction of wind-induced response of tall buildings, each with
WIND TUNNEL TESTING 49

its own strengths and limitations. These approaches, in order of their


complexity, are as follows:
• Nondirectional: This is a simple approach that is compatible with
code-mandated design wind speeds, but is the most conservative
approach, not taking into account any directionality.
• Sector method: This method introduces directionality, with the
highest directional wind speed(s) scaled to match code-mandated
values. As such, it is simple to understand and allows easy checking
of results but relies strongly on the judgment and expertise of practi-
tioners developing the directional model.
• Storm passage: This is a direct method of calculation using time
histories of storms, but it is computationally intensive and relies
heavily on the quality of the input wind data.
• Up-crossing: This was the first method used to provide a statistical fit
to the probability of occurrence of a load effect, but it includes some
approximations and is difficult to explain and check, owing to its
mathematical complexity.
• Multisector method: This is a more computationally intensive joint
probability approach that allows determination of the statistical
probability of a load effect using statistical fits to wind climate data
without any additional approximations.

Further details and comparisons of these techniques can be found in a


number of publications, including Davenport (1977), Lepage and Irwin
(1985), Gamble et al. (2001), Isyumov et al. (2003), Irwin et al. (2005),
Holmes (1990), and Holmes and Bekele (2015). There is not a complete
consensus in current wind engineering practice regarding the method used
to integrate measured wind-induced response with directional climate
effects, and this can vary geographically. For the design of a tall building
it is sensible to account for directional wind effects in a more robust way
than through the directional factor Kd in ASCE 7-16 (ASCE 2017).
Regardless of the approach taken, the wind consultant’s report should
include a discussion of the statistical method used to integrate wind climate
statistics with the wind effects measured in the wind tunnel and address the
limitations of the approach.

6.7 TYPICAL OUTPUTS

On completion of the wind tunnel tests, aerodynamic data are combined


with wind climate data to predict the overall wind-induced responses for
varying return periods, using approaches presented in Chapter 5.
Typical outputs from the predictions include maximum base moments
and shears for foundation design, effective static floor-by-floor loads at
50 DESIGN AND PERFORMANCE OF TALL BUILDINGS FOR WIND

appropriate MRIs for strength and serviceability considerations, and a rea-


sonable number of recommended load combinations to permit evaluation
of critical load effects in the building’s structure. An example is provided in
Figure 6-4, which illustrates the variation of base moments with wind direc-
tion and the sensitivity of loading to building frequency. The total number
of recommended load combinations may vary, but for a single tower it is
common to have between 10 and 24 possible loading scenarios to consider
for design.

Figure 6-4. Base moment response versus wind direction: sensitivity to frequency.
Source: Courtesy of CPP.
WIND TUNNEL TESTING 51

Component and resultant building accelerations are also predicted at


appropriate MRIs to permit evaluation of occupant comfort and tolerance
to building motions. The building accelerations should be compared to
applicable guidelines [e.g., ISO 10137 (2007)] and reviewed with the
building owner and other stakeholders to evaluate the serviceability perfor-
mance of the building for occupant comfort.

6.8 ADDITIONAL CONSIDERATIONS

In addition to the triggers for wind tunnel testing outlined earlier in this
chapter, there may be additional considerations that require a wind
consultant’s input to ensure a successful wind design for a tall building.

6.8.1 Shielding and Influence from Surrounding Buildings


The wind loads on a tall building in an urban setting may be influenced
by the interaction of wind with adjacent structures upstream of the build-
ing. Large buildings in close proximity to a study building may offer shel-
tering effects, which are beneficial. In other cases, the effects may be
adverse. The magnitude of wake effects created by an upstream building
is sensitive to the massing (shape, width, and height) of the adjacent
structure, as well as the spacing between the adjacent building and the
building under study. Because of the relative complexity of the issue, codes
of practice do not currently offer guidance on wake interference effects.
Wind tunnel tests are recommended to evaluate interference effects where
they are suspected to be adverse. If future development is known, this
should be included as an additional test configuration.

6.8.2 Design Evolution


Development of design may continue beyond the date of submission of
the wind consultant’s reports. Where revisions to the geometry of the build-
ing, effective sail area (changes in height or width), and/or dynamic struc-
tural properties has occurred, it is recommended that the revisions be
reviewed with the wind consultant to assess the impact of the changes.
If changes are deemed significant, additional testing may be required.

6.8.3 Minimum Thresholds


Lower limits on loading for MWFRS and pressures for components and
cladding should be considered. The following reflects the current state of
the art regarding minimum thresholds for wind loads.
When code-mandated minimum design wind speeds are reasonably
close (within approximately 5%) to actual predicted strength-level design
wind speeds, base overturning moments determined from wind tunnel
52 DESIGN AND PERFORMANCE OF TALL BUILDINGS FOR WIND

testing should be limited to not less than 80% of the design base overturning
moments determined in accordance with Part 1 of Chapter 27 of ASCE 7-16
(2017). If it can be shown that lower values are not a result of shielding from
specific influential nearby buildings, the lower limit can be reduced to 50%.
Additional testing with influential nearby buildings removed is usually
required to justify the reduced lower limit.
CHAPTER 7
DAMPING

7.1 OVERVIEW

Damping plays a key role in the wind-induced response of tall buildings.


The assumed level of damping in the design for strength and serviceability
is an important consideration for the structural engineer and wind tunnel
consultant. In general, three types of damping are considered for tall
buildings: inherent damping, aerodynamic damping, and supplemental
damping.

7.2 INHERENT DAMPING

Two primary sources contribute to inherent damping in tall buildings.


The first and most commonly known is the damping provided by the struc-
tural system itself. The second source is damping provided by secondary
nonstructural components of a building such as the exterior wall, interior
partitions, and finishes.
Damping provided by a structural system for wind-induced responses
varies depending on the type of structural system, the material composing
the system, and the amplitude of displacements under different recurrence
periods of considered wind events. Table 7-1 provides a summary of
common structural system critical damping ratios typically considered
for the design of tall buildings.
Assumed inherent damping levels for wind-induced response have been
traditionally lower than for seismic-induced responses because of assumed
lesser structural and nonstructural damage owing to wind loads. Building
height, aspect ratio, and structural system should be considered when selecting
53
54 DESIGN AND PERFORMANCE OF TALL BUILDINGS FOR WIND

Table 7-1. Recommended Critical Damping Ratios

Recommended critical damping ratios

Structural steel Reinforced concrete


MRI (years) systems (%) systems (%)

1–10 0.75–1.0 1.0–2.0

50–1,700 1.0–1.5 1.5–2.5

damping values to use for analysis. Damping ratios for hybrid or composite
systems may fall between the steel and concrete values noted previously.
Given the variability in reported damping levels in tall buildings, it would
be prudent for the structural designer and wind consultant to consider the
impact of damping assumptions parametrically as part of the design process.

7.3 AERODYNAMIC DAMPING

Aerodynamic damping of a tall building structure can be a significant


contributor to a slender tall building’s response to wind-induced forces
for both strength and serviceability considerations. This source of damping
is a result of the aerodynamic response of a tall building structure and is
influenced primarily by the building’s structural stiffness, generalized
mass, and building shape.
Although this damping is traditionally considered to be a positive
addition to a tall building response (i.e., increase in total damping), in
certain cases it can also negatively contribute to the effective damping of
a building, especially in the case of crosswind response close to the critical
reduced velocity. Estimates of aerodynamic damping can be made for sim-
ple shapes from published data (e.g., Steckley 1989). This aerodynamic
damping of tall buildings is most appropriately evaluated using aeroelastic
wind tunnel model testing as discussed in Chapter 6.

7.4 SUPPLEMENTAL DAMPING

With the relatively recent increase in the number of supertall towers, tall
building designs are now more commonly utilizing supplemental damping
systems to further control and optimize the structural system’s responses to
wind. With the goal of implementing additional effective damping to the
building’s response, supplemental damping can be provided by either
direct damping systems or, more commonly, indirect damping systems.
DAMPING 55

7.4.1 Direct Damping Systems


Where damping devices are installed and integrated directly into the
structural system of a tall building to provide supplemental damping, this
is considered a direct damping system. Although there are many propri-
etary products available, the vast majority of the devices available use vis-
cous or viscoelastic materials that provide additional damping when load
(or displacement) is applied to the device. Some examples of viscous or vis-
coelastic damping devices are piston-style fluid viscous damping devices
(Figure 7-1) and wall coupling/link beam polymer-joint damping devices
(Figure 7-2).
If direct damping devices are used, the design and analysis of the tall
building structural system must directly incorporate the dynamic behavior
of the devices being implemented because the devices are in the direct load
path of the lateral system.

Figure 7-1. Piston-style fluid damping device.


Source: Courtesy of Arup.

Figure 7-2. Wall coupling/link beam polymer-joint damping device.


Source: Courtesy of Kinetica.
56 DESIGN AND PERFORMANCE OF TALL BUILDINGS FOR WIND

7.4.2 Indirect Damping Systems


Indirect damping systems are the most common devices used to
provide supplemental damping for tall buildings and improve perfor-
mance. These devices are independent of the lateral load-resisting system.
They function by providing inertial mass that is tuned to oppose the
resonant wind-induced response of the structure. For a tall building,
the peak response of interest is typically at the top of the structure, which
corresponds to the most effective location to provide an indirect damping
system. The following different devices are used and are selected depend-
ing on the level of supplemental damping desired, as well as the required
effective mass:
• Tuned liquid column damper (Figure 7-3),
• Tuned sloshing tank damper (Figure 7-4),

Figure 7-3. Tuned liquid column damper.


Source: Courtesy of RWDI.

Figure 7-4. Tuned sloshing tank damper.


Source: Courtesy of RWDI.
DAMPING 57

• Tuned mass damper examples include a simple pendulum damper


(Figure 7-5) and compound pendulum damper (Figure 7-6), and
• Active mass damper (Figure 7-7).

Figure 7-5. Tuned mass damper: Simple pendulum damper.


Source: Courtesy of RWDI.

Figure 7-6. Tuned mass damper: Compound pendulum damper.


Source: Courtesy of RWDI.
58 DESIGN AND PERFORMANCE OF TALL BUILDINGS FOR WIND

Figure 7-7. Active mass damper.


Source: Courtesy of RWDI.

7.5 SUPPLEMENTAL DAMPING: STRENGTH AND


SERVICEABILITY

Traditionally, supplemental damping for wind resistance has been


implemented to improve the serviceability performance of the building.
However, in recent years, the tall building industry is taking steps to use
appropriate supplemental damping technology for strength design of the
primary structural system. Many factors are leading the industry to push
in this direction, including the high reliability of such damping devices,
and the consideration that supplemental damping systems are currently
used and prevalent in the strength design for seismic loading events in
the tall building industry.
If supplemental damping is used to reduce strength-level wind loads,
special consideration should be given to the appropriate strength-level
return periods to be considered with and without the additional damping,
such that temporary conditions before damper installation are sufficiently
covered in a tall building’s strength design. It is critical that the
chosen damping system is robust and redundant enough to achieve appro-
priate performance for strength-level design events. A robust long-term
inspection and maintenance program must also be implemented, in com-
pliance with manufacturers’ recommendations and design professionals’
requirements. It must be guaranteed that this system will perform and that
the required maintenance is conducted over the life of the building for a
supplemental damping system to be used for strength-level design
considerations.
CHAPTER 8
STRUCTURAL MODELING AND ANALYSIS

8.1 STRUCTURAL MODELING

The two primary objectives in the structural modeling and analysis of a


tall building for wind effects are to (1) estimate the structure’s stiffness and
mass as accurately as reasonably possible to allow reliable predictions of the
structure’s dynamic properties and movements, and to proportion the
structure with adequate stiffness for serviceability considerations; and
(2) evaluate the loads that are resisted by the elements that participate in
the lateral load–resisting system so that these elements may be propor-
tioned with adequate strength for ultimate load combinations.
For design purposes, it is neither realistic nor necessary for a structural
model to exactly replicate the building it is intended to represent.
A relatively simple model is often sufficient to achieve the design objectives
if the significant contributors to lateral stiffness and mass are considered
and appropriately included.
Numerous decisions and assumptions are necessary to construct and
analyze a structural model of a tall building, and these choices affect the
results of the analysis model. In turn, these results affect the structural
properties provided to the wind tunnel consultant, which will affect the
structural wind loads and accelerations, and how those loads are distrib-
uted to the individual elements composing the structural system.
This chapter discusses a variety of decisions that must be made in the
modeling and analysis phase of the design process, including
• Designation of lateral load–resisting elements,
• Mass assumptions,

59
60 DESIGN AND PERFORMANCE OF TALL BUILDINGS FOR WIND

• Consideration of P-Delta effects,


• Modeling assumptions for floor diaphragms,
• Foundation flexibility,
• Panel zone deformations at beam-column joints,
• Expected strength and modulus of elasticity of concrete materials, and
• Selection of stiffness reduction factors for reinforced concrete elements.

This chapter is not intended to provide a prescriptive set of recommen-


dations, but rather to highlight which important concepts should be
considered in the analysis of a tall building.

8.1.1 Strength-Level and Serviceability-Level Analysis


Because both strength-level and serviceability-level performance
objectives are considered, it is appropriate for distinct structural analysis
to be done for each of these levels. For reinforced concrete structures, differ-
ent stiffness assumptions may apply to each case. A strength-level analysis
alone may be sufficient in some cases in which building drift and acceler-
ations do not control the building’s performance. Multiple levels of service-
ability (e.g., 1-, 10-, 25-, and 50-year MRIs) could also be considered.

8.1.2 Primary Lateral Load–Resisting System and


Nonparticipating Elements
The engineer should identify the structural elements that are intended to
contribute to the primary lateral load–resisting system and those intended
to be gravity load–resisting elements only (nonparticipating). Appropriate
measures should be taken to either limit or account for the wind forces that
may develop in these nonparticipating elements owing to their relative
stiffness and/or connectivity.
Simplified structural models are usually preferred where possible;
however, the development of more powerful and user-friendly structural
analysis software has made it more practical for structural engineers to
add greater levels of detail to their analysis models. Although more detailed
structural models may be beneficial in some situations, unnecessary
increased detail can create more difficulty to clearly understand and verify
the appropriate load path for the distribution of the wind loads throughout
the structure. Further, modeling of all structural elements without
proper consideration may also result in the unintended consequence of
gravity-only elements attracting wind loads without proper design or
detailing.
The designer may choose to apply end releases or stiffness modifiers to
elements that are not intended to contribute to the resistance of lateral loads.
This is of primary importance for the strength-level model, in which it is
critical to ensure that no portion of the lateral load is neglected and that
STRUCTURAL MODELING AND ANALYSIS 61

the load is instead redistributed to the intended lateral load–resisting ele-


ments. For the service-level model, it may be preferable to omit these end
releases or stiffness modifiers to allow a more realistic estimate of stiffness
for serviceability considerations.
Both aforementioned approaches—using member end releases or stiff-
ness modifiers—implicitly rely on structural ductility, or at least concrete
cracking, to justify the redistribution of loads away from nonparticipating
elements. The engineer should identify conditions for which an unrealistic
amount of ductility is required. If this is the case, the originally “nonpartici-
pating” element should be considered part of the lateral load–resisting
system and should be designed and detailed for the induced lateral loads.
Nonstructural elements are always considered nonparticipating
elements. Cladding and partitions should be detailed based on the antici-
pated building movements, with joints large enough to avoid the participa-
tion of these elements in the resistance of lateral loads. If nonstructural
elements do attract lateral loads, the primary concern is that these elements
would be damaged, as discussed in Section 3.6. A secondary effect may be
that if the participation of nonstructural elements is significant (as has
been observed in some buildings with precast concrete cladding), the build-
ing may be noticeably stiffer than predicted by the analytical model.
Although such stiffness contributions from nonstructural elements may
benefit the final dynamic behavior of the building, these effects cannot
be reliably predicted in the design phase, and it would not be appropriate
to include any stiffness contribution from an element that is not designed
and detailed for the load it attracts.

8.1.3 Building Mass


The wind loads and accelerations determined by the wind tunnel
consultant are dependent on the modal information provided by the struc-
tural engineer. The self-weight of the structural elements of the building
contributes significantly to the overall mass, and typical assumptions for
the mass density of concrete may not be appropriate in cases in which high
strength and/or heavily reinforced concrete is used. When the structural
mass is determined automatically by the structural analysis software, care
should be taken to include mass for structural elements that were intention-
ally omitted for simplification of the analysis model.
Mass from the superimposed dead load (SDL) and live load (LL) should
be realistic rather than based on the values used for gravity system design,
which are typically conservatively set by the building code. The percentage
of SDL considered for building mass can vary depending on the building
type and the specifics of the finishes, but typical values range from 50% to
75% of the design loading. The LL measured in actual buildings has been
found to be substantially lower than the code-prescribed design LL, with
62 DESIGN AND PERFORMANCE OF TALL BUILDINGS FOR WIND

realistic ranges between 5% and 25%. Special consideration should be given


to ensure the mass at mechanical floors is appropriately estimated.
ASCE 7-16 (2017) Table C4.3-2 provides additional guidelines on expected
in-service live loads.

8.1.4 P-Delta (Second Order) Effects


P-Delta effects should always be considered, both for static and
dynamic/modal analysis. Most modern analysis software programs have
built-in capabilities to consider P-Delta effects. To implement these
capabilities, it is critical that the gravity loads are properly modeled such
that the software can accurately determine the P-Delta demands. The loads
applied should be those expected during the wind event being analyzed.
For serviceability models, the expected realistic dead and live loads should
be used. For strength models, the applied loads should be the factored
gravity loads from the ultimate load combination being considered.
For the case in which the analysis model for the lateral system omits
portions of the gravity system for simplicity, it is still necessary to include
the destabilizing effect of all gravity loads in the lateral analysis model.

8.1.5 Diaphragms
Diaphragms are structural elements that transfer loads from their points
of application to the lateral force–resisting elements or between different
elements of the lateral system. Diaphragm forces in tall buildings are
typically carried by the floor system. It is not always necessary to fully
model diaphragms in a lateral analysis for wind, but accurate assumptions
in model simplification are key to achieving an accurate distribution of lat-
eral loads in the building.
There are three main types of diaphragm modeling assumptions:
flexible, in which wind loads are distributed to lateral force–resisting
elements based solely on tributary wind sail area; rigid, in which all points
in the diaphragm are rigidly connected, resulting in wind loads distributed
based the lateral stiffness of the elements that are connected to the dia-
phragm; and semirigid, in which the stiffness of the diaphragm is explicitly
modeled, resulting in wind loads distributed based on the lateral stiffness of
the elements that are connected to the diaphragm and the stiffness of the
diaphragm itself. Flexible diaphragm assumptions are typically not
applicable in tall building analyses and are not addressed further in this
document.
Rigid diaphragm assumptions are reasonable for most above-grade
diaphragms in conventional tall buildings. As such, the diaphragm can
be modeled as a set of rigid body constraints applied to the vertical elements
at that level. This can save considerable computing demands on large build-
ings and is common in modern practice.
STRUCTURAL MODELING AND ANALYSIS 63

Semirigid diaphragms are necessary when modeling more complex


diaphragms or where transfer between lateral elements is expected.
Some examples of common configurations necessitating the explicit mod-
eling of semirigid diaphragms are as follows:
• Large reentrant corners, large openings, or openings located such that
local deformations can influence force distribution;
• Setbacks, transfers, or other major discontinuities of the lateral
force–resisting system;
• Outrigger and belt truss levels and other levels in which floors are
coupled by elements other than the continuous lateral force–resisting
system;
• Ground level floors where significant shear is transferred to basement
walls; and
• Any level where structural elements within the plane of the dia-
phragm carry significant axial forces, such as braced frame chords,
truss chords, or in-plane bracing.

Diaphragms should be designed to carry the loads determined by the


analyses. When the diaphragm is not explicitly modeled (i.e., a rigid
diaphragm assumption is used), a simplified analysis can be performed
with loads applied based on tributary area or another rational distribution
method.

8.1.6 Foundation Flexibility


Before the advent of modern computing, most tall buildings were
designed assuming pinned bases at the foundation level, and this is usually
understood to have resulted in satisfactory designs. With the increase in
computing power and advancements in analysis software, it is now more
practical to model the flexibility of the foundations of a building. This can be
accomplished by assigning spring stiffnesses at the base of the structure, or
even by explicitly modeling foundation elements and soil layers within the
structural analysis model.
Modeling the flexibility of the foundations is not usually considered in
the calculation of the building dynamic properties provided to the wind
tunnel consultant. When foundation flexibility is included, the amount of
inherent damping used in the calculation of wind response should be
increased to account for the additional damping contributed by the soil–
structure interaction. Ultimately, engineering judgment should be used
to determine when the foundation flexibility should be modeled. In some
cases, doing so may be necessary to adequately capture the load distribu-
tion in the structure for strength design, even if foundation flexibility is
neglected for the computation of dynamic properties.
64 DESIGN AND PERFORMANCE OF TALL BUILDINGS FOR WIND

8.1.7 Panel Zone Deformations


Panel zones are the regions of beam-column intersection in moment
frames (Figure 8-1). At these locations, flexural deformations are limited
by the depth of the connecting member, but the concentrated moments
on each side of the column create a localized shear within the panel zone.
Because the resulting shear deformations within the panel zones may be
significant, panel zones should not be modeled as rigid.
There are several options for considering panel zone deformations. The
simplest method is to use a centerline model without explicit consideration
of panel zones. The resulting racking story drifts in centerline models are
typically larger than those found when incorporating elastic panel zone
deformations and beam/column offsets, because the overestimation of
flexural deformation will usually exceed the underestimation of shear
deformation within the panel zone. However, there are some cases in which
a centerline model will overestimate stiffness, particularly if there is shear
yielding within the panel zone.
The more accurate approach is to model the panel zones, which will
usually result in improved stiffness estimates compared to a centerline
model. This will also result in more accurate estimates of building dynamic
properties, and thus a more accurate wind load determination.
Published methods for modeling the panel zones include the Krawinkler
joint model and the scissors joint model (Charney and Marshall 2006).
Alternatively, the panel zones may be studied using a series of subassembly
finite-element models, and stiffness modifiers may be calculated and
applied to centerline models.

Figure 8-1. Panel zone shear at moment frame beam-column intersection.


STRUCTURAL MODELING AND ANALYSIS 65

8.2 SPECIAL CONSIDERATIONS FOR REINFORCED


CONCRETE STRUCTURES

In contrast to structural steel, which demonstrates linear-elastic behavior


below the yield stress with a well-defined modulus of elasticity, concrete
behavior is more complicated in several ways. Concrete demonstrates:
• Nonlinear stress-strain relationship;
• Differing behavior in compression and tension, with tensile cracking
occurring at stresses far below the compressive strength;
• Significant changes in stiffness based on degree of cracking; and
• Inherent variability in concrete material properties and the expected
strength and stiffness.
These issues merit special consideration during the design of reinforced
concrete structures and are addressed in the following sections.

8.2.1 Expected Strength and Modulus of Elasticity of Concrete


Materials
In consideration of the statistical variability of concrete strength, ACI
301-16 (ACI 2016) requires that the average strength of a concrete mix is
greater than the specified strength by some margin, which may be deter-
mined based on the sample standard deviation. This ensures a suitably
low probability that a given sample will have a strength less than the
specified value, but it also typically results in in situ strength greater than
the specified value. This also implies that most of the concrete placed in the
field will have a higher than anticipated corresponding modulus of elastic-
ity (MOE).
Because the lateral stiffness of a tall building is most closely related to the
average MOE of the concrete, rather than the minimum value, it is appro-
priate for analysis to use the average MOE for a given concrete mix. ACI
363R-10 (ACI 2010) summarizes empirical data, which suggest that the
standard ACI 318-14 (ACI 2014) equation for the modulus of elasticity
for normal density concrete is usually an appropriate predictor of the
average MOE for high-strength concretes used in tall buildings. Refer to
ACI 363R-10 for further discussion, including alternative equations for
MOE that may be more accurate if information about the mix design, coarse
aggregates, and curing conditions are known.
When data on expected average MOE for a given mix design is available,
this may be used as a more accurate alternative to calculating the modulus
based on the compressive strength. When stiffness is a critical design
consideration, the designer should specify the required MOE, and the
contractor should evaluate the MOE through direct measurement and pre-
construction testing if representative mix design test data is not available.
66 DESIGN AND PERFORMANCE OF TALL BUILDINGS FOR WIND

The testing should demonstrate that the average MOE of the samples meets
or exceeds the specified value; it is not necessary to require that all samples
exceed the specified value.
It is reasonable to allow the desired MOE to be achieved later than the
standard 28-day or 56-day ages used for strength testing, because this will
be more representative of the stiffness available during the service life of the
building. For practical reasons, it still may be preferable to test 28-day or
56-day samples rather than samples that are significantly older. In this case,
the designer may perform serviceability checks using an extrapolated MOE
for more mature concrete if realistic concrete strength-versus-time data is
available for a representative mix.

8.2.2 Stiffness Modifiers and Behavior of Cracked Reinforced


Concrete Structures
For buildings with reinforced concrete lateral load–resisting systems, the
structure’s lateral stiffness is usually sensitive to the extent of concrete
cracking, so it is essential to account for cracking in the lateral analysis
of any reinforced concrete structure. Cracking may either be modeled
explicitly with material nonlinearity of concrete or may be modeled
implicitly by applying stiffness reduction factors.
An implicit approach using stiffness reduction factors is common in
practice for wind design because it allows for an elastic analysis. The
discussion in this section focuses on this approach.
When stiffness reduction factors are used, the goal should be to calibrate
the reduction factors such that a linear analysis at a given load level results
in displacements similar to those obtained from a more rigorous nonlinear
material analysis or from field measurements of the completed building.
Realistically, however, there will inevitably be differences in predicted
versus actual stiffness. In acknowledgment of this uncertainty, it may be
prudent to consider a range of stiffness modifiers, especially if wind effects
are highly sensitive to the building period (see Figure 6-4 and Section 6.7 for
further discussion).
In general, different stiffness reduction factors should be used for differ-
ent magnitudes of load. Because the typical reinforced concrete force-
displacement curve continues to flatten as the load increases and cracking
becomes more extensive, the stiffness reduction factors are usually lower at
ultimate load levels compared to service load levels.
There are two different approaches for implementation of stiffness
reduction factors to account for cracking: (1) the simplified method in which
stiffness modifiers are selected prior to the analysis, with little or no itera-
tion required; and (2) the detailed method in which stiffness modifiers are
selected in consideration of reinforcement and load in the element. In the
detailed method, iteration is required.
STRUCTURAL MODELING AND ANALYSIS 67

8.2.3 Simplified Method for Selecting Stiffness Modifiers


The first approach is to select modifiers based on the type of structural
element (wall, column, beam, and slab, among others) and level of loading
(serviceability or strength) and use the same modifier for elements of the
same type throughout the height of the building. If this method is used,
Table 8-1 shows an example set of reduction factors that are representative
of typical practice.
The values shown in Table 8-1 are intended to be an example of stiffness
assumptions that may be appropriate for a typical structure, but given the
inherent limitations of the simplified method, the tabulated values should
not be considered mandatory. Alternate values may be more appropriate
for a given structure. Determination of stiffness factors using rational analy-
sis, such as the methods described in Section 8.2.4, will usually result in
greater accuracy even if the detailed approach is not fully adopted for all
elements.
Implicit in the values from Table 8-1 is the expectation, in advance of
the analysis, that some element types will be cracked, and some elements
types will be uncracked. In general, uncracked elements have a modifier
of 1.0 on the gross section properties—although modifiers greater than
1.0 can be justified if the reinforcement is considered in the calculation
of transformed section properties—and cracked element types have
modifiers less than 1.0.
Shear walls are given special consideration in that different values are
listed for uncracked shear walls and cracked shear walls. This requires an
initial analysis in which the stresses in the walls are checked under the
envelope of applicable load combinations, and if the tension stress exceeds
qffiffiffiffi
0
the rupture stress (recommended to be taken as 6 f c for walls), then the
walls are considered cracked. The extent and magnitude of the tension
stresses should be considered when selecting a value within the suggested
range. If there is only a small localized area of wall where the stress exceeds
the rupture stress, then the upper bound modifier may be appropriate,
whereas a lower bound modifier would be appropriate for a wall with
tension stress greater than the rupture stress over the entire cross section.
As stated in Notes f and g of Table 8-1, a similar approach may also be
used to determine whether other element types would be considered
cracked or uncracked for a given level of loading.

8.2.4 Detailed Method for Selecting Stiffness Modifiers


As an alternative to the simplified method described in Section 8.2.3, the
designer may choose to use a more detailed approach and select unique
stiffness reduction factors for individual elements. This approach
68 DESIGN AND PERFORMANCE OF TALL BUILDINGS FOR WIND

Table 8-1. Example Reduction Factors for Reinforced Concrete Elements

Serviceability analysis Strength analysis


Wind load MRI <100 yearsa >700 yearsb

Column axialc 1.0 Ag 1.0 Ag

Column flexure 1.0 Ig 0.7 Ig


RC frame beams 0.5 Ig 0.35 Ig

PT frame beams 0.7 Ig 0.5 Ig

RC slabs 0.35 Ig 0.25 Ig

PT slabs 0.5 Ig 0.35 Ig

Uncracked shear walls 1.0 Ig 0.875 Ige


axial-flexured

Cracked shear walls 0.6-0.9 Ig 0.5-0.8 Ig


axial-flexured
Link beam flexuref 0.5 to 0.7 Ig 0.3 to 0.6 Ig

Link beam shearg 0.15 to 1.0 Ag 0.1 to 0.7 Ag


a
Typically, 1, 10, 25, or 50 years. Corresponds to applicable performance objective.
See commentary to Appendix C of ASCE 7-16 (2017).
b
Per governing code.
c
Columns resisting overturning forces and under axial tension should use shear wall values for
axial stiffness modification.
d
Walls with axial tensile stress greater than the modulus of rupture are considered cracked.
qffiffiffiffi
0
A modulus of rupture of 6 f c is suggested for walls. Uniform cracking modifier is applied
to vertical membrane stiffness of all shear wall elements in a lateral system. Selection of
cracking modifier within suggested range should be based on the extent of cracking observed.
e
Corresponds to stability stiffness modifier fk recommended by R6.7.1.1 of ACI 318-14
(ACI 2014).
f
Link beams with flexural stress greater than the modulus of rupture are considered cracked in
flexure. Link beams that are uncracked in flexure may be considered with flexural stiffness
based on 1.0 Ig. For link beams that are cracked in flexure, lower bound values are
appropriate for link beams with light flexural reinforcement, and upper bound values are
appropriate for link beams with heavy flexural reinforcement.
g
Link beams with shear stress greater than the code-based strength for concrete alone are
considered cracked in shear. Link beams that are uncracked in shear may be considered
with shear stiffness based on 1.0 Ag. For link beams that are cracked in shear, ACI SP-240-5
(2006) reports modifiers between 0.1 and 0.4, with lower bound values appropriate for link
beams with light shear reinforcement and upper bound values appropriate for link beams
with heavy shear reinforcement.
Note: For uncracked elements of all types, transformed section properties (including stiffness
contribution of reinforcement) may be used in lieu of gross section properties.
STRUCTURAL MODELING AND ANALYSIS 69

recognizes that two individual elements of the same type will, in general,
have differing
• Cross sections,
• Loads and stresses,
• Extent of cracking, and
• Reinforcement.

Any one of these factors, or any combination of them, would mean


that the two individual elements of the same type would have different
force-displacement behavior, which implies that different stiffness
modifiers are appropriate for each.
Calculation of stiffness modifiers using the detailed method should be
based on the actual load and reinforcement in the element and should
approximate the secant stiffness based on that load and the expected
force-displacement behavior (Figure 8-2). Different stiffness modifiers
would usually be expected for the same element under service loads and
ultimate loads.
As an example of the detailed approach, the flexural stiffness of a
reinforced concrete beam may be calculated in consideration of its moment
and reinforcement using the well-known Branson equation from ACI 318 as
follows:

 3    
Mcr Mcr 3
Ie ¼ Ig + 1 − I cr
Ma Ma

Figure 8-2. Secant stiffness approximation.


70 DESIGN AND PERFORMANCE OF TALL BUILDINGS FOR WIND

The following equations for the shear stiffness of a reinforced concrete


beam, in consideration of shear cracking, are recommended in ACI
Committee 375’s SP-240-5 (2006) based on the work of Dilger and
Abele (1974):

Gbw d
GAv ¼ Gross GAv ¼ for V ≤ V c
1.2
ρv
GAv cr ¼ Es bw d
1 + 4nρv − VVc

where
Vc = Concrete shear strength
V = Applied shear
ρv = Shear reinforcing volume ratio = bAwvs
Av = Shear reinforcing area
s = Spacing of shear reinforcing
n = Modular ratio = EEcs
Es = Modulus of elasticity of reinforcing steel
Ec = Modulus of elasticity of concrete
bw = Width of rectangular beam
d = Effective depth of beam
For tall reinforced concrete buildings, shear walls often provide most of
the structure’s lateral stiffness, and the lateral stiffness of the shear walls is
sensitive to axial-flexural cracking. It may be prudent to scrutinize the
stiffness of these elements more closely by using the detailed approach,
while using the simplified approach for the other, less significant
contributors to the lateral system.
One approach for evaluating the stiffness of a cracked reinforced concrete
wall is presented by Rahimian (2011) and includes consideration of tension
stiffening behavior, that is, the stiffness provided by the concrete in between
crack locations. Using Rahimian’s explicit method, the flexural cracking modi-
fier for any wall element may be calculated based on the concrete stress-strain
curve in tension, the wall’s vertical reinforcement ratio, and the wall tensile
stress observed from a linear analysis. This calculation may be performed
for each cracked wall element, and a unique cracking modifier may be
assigned to each. Iteration is required, because the extent of cracking will usu-
ally increase once the stiffnesses of the initial cracked elements are reduced.
For greatest accuracy, the described iterative procedure would be
performed independently on different versions of the model—one for each
load combination being considered. Alternatively, the wall stress used in
the calculation may be conservatively based on the envelope of load
combinations.
CHAPTER 9
WIND OPTIMIZATION PROGRAM

9.1 INTRODUCTION

A building’s aerodynamic shape plays a significant role in the design of


tall and supertall buildings. The goal of the aerodynamic optimization
process is to determine a building geometry that provides near-optimal
wind response, while also meeting all other design requirements, including
architectural constraints, cost targets, and construction considerations.
Aerodynamic optimization may therefore be classified into two categories:
aerodynamic modifications and aerodynamic design.
Aerodynamic modifications are relatively minor changes to building
shape or massing owing to architectural constraints that limit the modifi-
cations to enhance wind performance. Hence, corner treatments, such as
chamfering, slotting, and roundness are common approaches in this
category. Aerodynamic design, on the other hand, is an approach that is
integrated with the architectural design from the early stages of the project.
The building shaping, massing orientation, and geometry are permitted to
be influenced by results of wind studies. A series of wind tunnel tests are
usually required to achieve the required number of design iterations.

9.2 BUILDING ORIENTATION

The first four links in the Davenport wind loading chain (the commonly
used method to determine wind-induced responses of buildings) are wind
climate, influence of terrain, aerodynamic effects, and dynamic effects
(Davenport 1977, 1982). Once a detailed wind climate and wind tunnel
study have been completed, each of these parameters will have been
71
72 DESIGN AND PERFORMANCE OF TALL BUILDINGS FOR WIND

determined on a directional basis. Optimizing a building’s orientation


aligns these directional distributions such that the wind load effect of
interest is minimized.
When considering wind optimization through-building orientation, it is
recommended that the combined effects of wind climate and terrain be
considered. Combining the wind climate model, which is natively in a refer-
ence open terrain, with the terrain and topography model will produce
more valuable directional insight for the design team.
Changing the orientation of the building at its base requires a flexible site
location, where the building footprint is relatively unconstrained by
adjacent streets and buildings. Additional flexibility is typically available
at higher elevations, where a twist or change in orientation may be intro-
duced. This can allow for the upper portions of a tall building to be more
favorably oriented with respect to the directionality of the wind climate, in
addition to the potential reduction in crosswind loading.
The magnitude of reductions available from building orientation
optimization depend on the directionality of the wind climate, wind load
effects, and the flexibility of the site location. In some circumstances, this
optimization could yield reductions upward of 25% for base building
overturning moments.

9.3 BUILDING GEOMETRY

The architectural form of tall slender buildings tends to be heavily


influenced by the impact of wind, which is dependent on the geometry
(i.e., shape) of the building itself. The dynamic wind response of tall struc-
tures is governed by a number of factors including shape, stiffness, mass,
and damping. It is not uncommon for modern tall slender towers to suffer
from dynamic vortex-shedding issues at relatively frequent recurrence
intervals. The vortex-shedding characteristics of a tower are dependent
on the basic tower shape via the nondimensional quantity known as the
Strouhal number.
Aerodynamic modifications of the tower shape can help to control
dynamic motions and reduce the risk of excessive dynamic wind response.
Potential modifications include recessed, slotted, or chamfered corners,
horizontal and vertical through-building openings, porous tops, tapering
footprints, and staggered (dropping off) corners, all of which have been
shown to significantly reduce the wind-induced loads and responses in tall
buildings. Figure 9-1 shows examples of corner modifications.
Chamfers, twisting, and tapering of the building reduce the potential for
a tower to exhibit dynamic instability, as illustrated in Tanaka et al. (2013).
Aerodynamic modifications achieve this most effectively by reducing the
correlation of the vortex shedding along the height of the tower.
WIND OPTIMIZATION PROGRAM 73

Figure 9-1. Aerodynamic modifications to a square building shape.


Source: Adapted by Kijewski-Correa from Kareem et al. (1999).

Figure 9-2. Rate of twist shape optimization (the Chicago Spire).


Source: Courtesy of TT/RWDI.

In general, modifications to the building corners such as slotted or


chamfered corners typically need to be greater than approximately 5% to
10% of the building width to be beneficial. Modifications that increase
the projected area or the breadth of the building may improve crosswind
responses, but can result in increased along-wind loads and responses.
Similarly, twisting or tapering need to be significant to be substantially
beneficial (Figure 9-2). Aerodynamic modifications are usually most
effective when applied over the top third of the building height. The
approximate effectiveness of some of the measures, in terms of the reduc-
tion of the dynamic loads or accelerations from a standard baseline building
shape with a square footprint, is given in Table 9-1.

9.4 HOLISTIC OPTIMIZATION

It is relatively straightforward for wind engineers to suggest


aerodynamic improvements to buildings to reduce loads and responses;
74 DESIGN AND PERFORMANCE OF TALL BUILDINGS FOR WIND

Table 9-1. Approximate Effectiveness of Aerodynamic Modifications

Aerodynamic Dynamic load


modification reduction (%) Dependent on

Baseline square 0 —

Rounded corners 10–25 Shape and size of rounding


Chamfer 10–25 Shape and size of chamfer

Taper 0–25 Degree of taper

Twist 0–50 Degree of twist

Combination of 25–75 Extent of modifications


modifications

however, to be an effective contributor to design, these modifications need


to be considered holistically within the total design environment.
The first consideration is site and building design and aesthetic
constraints. The second constraint is the program and floor area require-
ments for the project. The third constraint for many locations is maximum
height restrictions. Other considerations include the efficiency of floor plate
and view corridors.
In most cases, optimization is conducted to control crosswind responses,
which requires a combination of aerodynamic form and structural
optimization. Control of along-wind responses is more dependent on aero-
dynamic shape and orientation relative to critical wind directions with the
structural properties typically being less significant.
If a building is already at the maximum permitted height, this limits the
potential for aerodynamic shaping, which usually implies a reduction in
usable space on modified floors, unless there is room to increase the floor
plate size. In very dense cities, this is often not possible. In contrast, on
larger open sites increasing plan dimensions may be possible.
The effects of aerodynamic optimization also need to be considered
in relation to the effects on the structural performance of revised architec-
tural forms, because changes in dynamic properties also affect the wind
loads. Such changes in properties must be considered with regard to
the critical design responses, whether these are accelerations, serviceability
deflections, or strength-level design loads (Denoon et al. 2012; Tse et al.
2006a, b).
WIND OPTIMIZATION PROGRAM 75

Usually, aerodynamic optimization occurs after a set of initial wind


tunnel tests have been conducted to establish a baseline against which
alternative designs can be compared. This allows for the identification of
the responses that are critical for design and the most appropriate form of
optimization, whether aerodynamic, structural, or a combination of both, to
then be determined. Aerodynamic shaping options that increase the build-
ing height have the potential to increase the along-wind responses, but if
these do not control the loads or accelerations, this may be acceptable. In
addition, the shaping may result in sufficiently improved response such
that the need for supplemental damping can be reduced or altogether elim-
inated.
CHAPTER 10
CONCLUDING REMARKS

10.1 DESIGN VALIDATION

As discussed in Chapter 2, the design process for tall buildings starts


with the establishment of performance objectives and a preliminary
structural design. This is followed by a wind climate assessment and wind
tunnel study that provides more accurate wind loads for use in design and
evaluates the building accelerations and movement under wind loads. The
results of the wind tunnel study are then used to develop a final design that
meets the performance objectives.
When there are significant changes in building geometry, orientation,
surroundings, or dynamic properties, it is recommended that the wind
tunnel consultant review the changes and advise the design team on the
implications of the changes on wind loads and building acceleration.
Often, the wind tunnel consultant can extrapolate new building wind loads
and dynamic response based on the original wind tunnel test. However,
when the building geometry or environment changes significantly,
retesting in the wind tunnel may be required. The final wind tunnel results
are used for completing the structural design and producing final design
documentation to be used for construction.

10.2 PEER REVIEW

For tall building design, some jurisdictions may require a peer review of
the structural design as part of the approvals process when certain
conditions related to height, area, or slenderness ratio are triggered.
However, even if a peer review is not required by code or the local
77
78 DESIGN AND PERFORMANCE OF TALL BUILDINGS FOR WIND

jurisdiction, owners often consider engaging a third-party peer reviewer to


review the wind design of the primary structure for additional assurance of
the quality of the design and for potential value engineering benefits.
Owners may consider a peer review when the building height is greater
than 400 ft (approximately 120 m) or for slender or buildings with otherwise
unique attributes such as unusual or complex geometries or foundations.
An independent peer review brings a different perspective and often helps
identify aspects of a building design that warrant special attention.
If a code peer review is required, the code or local jurisdiction likely
prescribes the requirements for the review scope and process. For owner-
initiated peer reviews, the peer review may occur near the completion of the
design, when the owner seeks to verify the safety of the design. If the owner
is looking for value engineering benefits, a multistage peer review may be
implemented to obtain input at various stages of design. This can help
identify and resolve potential issues as well as allow for value engineering
earlier in the design process when it is easier to incorporate design changes.
It is recommended that the structural peer review be performed by a
qualified independent structural engineer with experience in tall building
design. The structural engineer of record should provide the peer reviewer
with the drawings and specifications and structural reports, as well as the
wind tunnel and geotechnical reports. The owner may also request that the
engineer provide the structural analysis model to the peer reviewer or ask
that the peer reviewer develop an independent analysis.
The peer reviewer should provide written comments to the structural
engineer, owner, and, if required, to the building department having
jurisdiction. The structural engineer would then provide written
responses to the review comments. Sometimes, multiple rounds of
comment/response may be required to resolve issues.

10.3 CONCURRENT RESEARCH AND FUTURE DIRECTIONS

10.3.1 Monitoring
Monitoring of building properties and responses can be used both
during and after construction to validate and/or refine wind tunnel predic-
tions. Measurements of natural frequencies and (estimates of) inherent
structural damping can be used to provide updated predictions of building
responses, particularly to help tune any supplementary damping require-
ments. Longer-term measurements of building accelerations can be used
during occupancy to ensure that the building is performing as expected
and to address any serviceability issues that might be associated with
building movement. More common monitoring of field performance of
completed buildings, with results published in the public domain, can only
CONCLUDING REMARKS 79

assist advances in structural engineering to increase future design efficiency


and building performance.

10.3.2 Performance-Based Design


Performance-based design (PBD) procedures have been developed in
recent years for use in seismic design. Performance-based wind design is
at an early stage of development and research, with multiple teams work-
ing on the advancement of appropriate methodologies. One of the first out-
comes from this is the ASCE/SEI Prestandard for Performance-Based Wind
Design (2019). This Prestandard describes a basis for how to use PBD for
both structural and envelope design for buildings. It is applicable for all
buildings and provides guidance on performance goals and minimum
requirements that must be met in design to demonstrate that these have
been met.

10.3.3 Computational Wind Engineering


Computational wind engineering, commonly using computational fluid
dynamics, is sometimes promoted as an alternative to wind tunnel testing
to determine wind effects on buildings and structures as is done in the
aerospace and automobile industries. However, ensuring the accuracy
and reliability of CWE procedures requires highly experienced analysts
with wind engineering backgrounds. Loads and responses based on
CWE need to be carefully validated, and the reliability of CWE remains
unproven for structural engineering applications. At the time of publica-
tion, there is no ASCE-accepted standard that allows the use of CWE for
structural design, and there are limited validated studies that show that
the use of CWE can meet the structural reliability requirements of
ASCE 7. CWE can, however, be used for a range of other studies of tall
building design related to serviceability issues such as pedestrian wind
comfort or the investigation of air flows within the building.

10.3.4 High-Performance and New Materials


Material advancements in the use of now readily available ductile
high-strength steels, as well as high-strength and high-modulus concretes,
will drive industry and design of tall buildings. Their constructability
advancements will also be critical in their further adoption into the tall
building design industry.
New lightweight materials, such as engineered mass-timber may also be
considered in the design of tall buildings and thus warrant consideration.
Special attention and research should be dedicated to investigating
mass-timber tall buildings with lower-mass aerodynamic behavior,
increased inherent damping, and their system durability and robustness.
80 DESIGN AND PERFORMANCE OF TALL BUILDINGS FOR WIND

10.4 CLOSING REMARKS

As design and construction techniques rapidly evolve, so will the


building industry’s ability to fulfill societal ambitions for taller and more
complex structures. The consideration of wind effects on tall buildings will
become increasingly important as new structures continue to push the
bounds of what has been previously designed. This Manual of Practice
provides the framework with which tall buildings can be evaluated for
wind effects. Given the continuous innovation in the field of tall building
design, it is critical that structural and wind engineers keep current on new
developments.
REFERENCES

ACI (American Concrete Institute). 2006. Performance-based design of concrete


building for wind loads. ACI SP-240. Farmington Hills, MI: ACI.
ACI. 2010. Report on high-strength concrete. ACI 363R-10. Farmington Hills,
MI: ACI.
ACI. 2014. Building code requirements for structural concrete. ACI 318-14.
Farmington Hills, MI: ACI.
ACI. 2016. Specifications for structural concrete. ACI 301-16. Farmington Hills,
MI: ACI.
AIJ (Architectural Institute of Japan). 2004. Guidelines for the evaluation of
habitability to building vibration. AIJES-V001-2004. Tokyo: AIJ.
AISC (American Institute of Steel Construction). 2016. Specification for
structural steel buildings. AISC 360-16. Chicago: AISC.
ANSI (American National Standards Institute). 1972. Building code require-
ments for minimum design loads in buildings and other structures. ANSI
A58.1-1972. New York: ANSI.
ANSI (American National Standards Institute)/ASCE. 1988. Minimum
design loads for buildings and other structures. ANSI/ASCE 7-88. Reston,
VA: ASCE.
ASCE. 1999. Wind tunnel studies of buildings and structures. ASCE MOP 67.
Reston, VA: ASCE.
ASCE. 2012. Wind tunnel testing for buildings and other structures. ASCE/SEI
49-12. Reston, VA: ASCE.
ASCE. 2017. Minimum design loads and associated criteria for buildings and
other structures. ASCE/SEI 7-16. Reston, VA: ASCE.
ASCE. 2019. Prestandard for performance-based wind design. Edited by Scott,
D. R. Reston, VA: ASCE.
AS/NZS (Standards Australia Limited/Standards New Zealand). 2011.
Structural design actions, Part 2: Wind actions. AS/NZS 1170.2:2011.
Wellington, NZ: AS/NZS.

81
82 REFERENCES

Burton, M., K. C. S. Kwok, and A. Abdelrazaq. 2015. “Wind-induced


motion of tall buildings: Designing for occupant comfort.” Int. J. High-
Rise Build. 4 (1): 1–8.
Charney, F. A. 1990. “Wind drift serviceability limit state design of
multistory buildings.” Wind Eng. Ind. Aerodyn. 36 (1–3): 203–212.
Charney, F. A., and J. Marshall. 2006. “A comparison of the Krawinkler and
Scissors models for including beam-column joint deformations in the
analysis of moment-resisting steel frame.” Eng. J. 43 (1): 31–48, AISC.
Davenport, A. G. 1977. “The prediction of risk under wind loading.”
In Proc., 2nd Int. Conf. on Structural Safety and Reliability. Munich, Germany.
Davenport, A. G. 1982. “The interaction of wind and structures.” In
Engineering meteorology, edited by J. Plate. Amsterdam, Netherlands:
Elsevier.
Denoon, R., K. Strobel, and D. Scott. 2012. “Challenging paradigms in wind
engineering design of tall buildings.” In Proc., 9th World Congress of
Council on Tall Buildings and Urban Habitat. Shanghai, China.
Dilger, W. H., and G. Abele. 1974. “Initial and time dependent shear deflec-
tions of reinforced concrete T-beams.” In Deflections of concrete structures.
487–514. Farmington Hills, MI: ACI.
FEMA. 2018. Seismic performance assessment of buildings. FEMA P-58.
Washington, DC: FEMA.
Gamble, S. L., R. J. Miltenburg, M. D. Cicci, and M. Accardo. 2001.
“Prediction of local exterior wind pressures from wind tunnel studies
using a time history analysis approach.” In Proc., 1st Americas Conf. on
Wind Engineering. June 3-6, 2001. Clemson, SC.
Griffis, L. 1993. “Serviceability limit states under wind load.” Eng. J. 30: 1–16.
Harris, R. I., and D. M. Deaves. 1980. “The structure of strong winds.”
In Proc., CIRIA Conf. London. .
Holmes, J. D. 1990. “Directional effects on extreme wind loads.” In Civil engi-
neering transactions, 45–50. Sydney, Australia: Institution of Engineers.
Holmes, J. D., and S. A. Bekele. 2015. “Directionality and wind induced
response–Calculation by sector methods.” In Proc., 14th Int. Conf. on
Wind Engineering (ICWE14). Porto Alegre, Brazil.
Irwin, P., R. Denoon, and D. Scott. 2013. Wind tunnel testing of high-rise
buildings. London: Routledge.
Irwin, P., J. Garber, and E. Ho. 2005. “Integration of wind tunnel data with
full scale wind climate.” In Proc., 10th Americas Conf. on Wind Engineering.
May 31-June 4, 2005. Baton Rouge, LA.
Irwin, P., and B. Myslimaj. 2008. “Practical experience with wind-tunnel
predicted tall building motions.” In Vol. 17 of Proc., IABSE Congress
Report, International Association for Bridge and Structural
Engineering (IABSE), Chicago: Zurich. 296–305.
REFERENCES 83

ISO. 1984. Guidelines for the evaluation of the response of occupants of fixed
structures, especially buildings and off-shore structures, to low-frequency
horizontal motion (0, 063 to 1 Hz). ISO 6897:1984. Geneva: ISO.
ISO. 2007. Bases for design of structures–Serviceability of buildings and walkways
against vibrations. ISO 10137:2007. Geneva: ISO.
Isyumov, N., M. J. Mikitiuk, P. C. Case, G. R. Lythe, and A. Welburn. 2003.
“Predictions of wind loads and responses from simulated tropical storm
passages.” In Proc., 10th Int. Conf. on Wind Engineering. Copenhagen,
Denmark.
Kareem, A., T. Kijewski, and Y. Tamura. 1999. “Mitigation of motions of tall
buildings with specific examples of recent applications.” Wind Struct.
2 (3): 201–251. 10.12989/was.1999.2.3.20110.12989/was.1999.2.3.201.
Lepage, M. F., and P. A. Irwin. 1985. “A technique for combining historical
wind data with wind tunnel test to predict extreme loads.” In Proc., 5th
US National Conf. on Wind Engineering. Lubbock, TX.
Li, Q. S., J. Q. Fang, A. P. Jeary, and C. K. Wong. 1985. “Full scale measure-
ments of wind effects on tall buildings.” J. Wind Eng. Ind. Aerodyn. 74–76:
741–750.
NRC (National Resource Council). 2015. National building code of Canada.
Ottawa, ON: NRC.
Rahimian, A. 2011. “Lateral stiffness of concrete shear walls for tall
buildings.” Struct. J. 108 (6): 755–765.
Riad, J. 2016. “Conceptual high-rise design—A design tool combining
stakeholders and demands with design.” M.S. thesis, Dept. of Applied
Mechanics, Division of Material and Computational Mechanics,
Chalmers Univ. of Technology.
Simiu, E., and D. Yeo. 2019. Wind effects on structures: Modern structural
design for wind. 4th ed. Oxford, UK: Wiley-Blackwell.
Steckley, A. 1989. “Motion-induced wind forces on chimneys and tall
buildings.” Ph.D. thesis, Faculty of Graduate Studies, Univ. of
Western Ontario.
Tanaka, H., Y. Tamura, K. Ohtake, M. Nakai, Y. C. Kim, and E. K. Bandi.
2013. “Aerodynamic and flow characteristics of tall buildings with vari-
ous unconventional configurations.” Int. J. High-Rise Build. 2 (3): 213–228.
Tse, K. T., K. C. S. Kwok, P. A. Hitchcock, C. M. Chan, and R. O. Denoon.
2006a. “Building economics of wind-engineered tall structures—Part 1:
Empirical response formulae of corner-modified buildings.” In Proc.,
12th Australasian Wind Engineering Society Workshop. Queenstown, NZ.
Tse, K. T., K. C. S. Kwok, P. A. Hitchcock, C. M. Chan, and R. O. Denoon.
2006b. “Building economics of wind-engineered tall structures—Part 2:
Construction cost and returns on investment.” In Proc., 12th Australasian
Wind Engineering Society Workshop. Queenstown, NZ.
USG Corporation. 2014. The gypsum construction handbook. 7th ed. Hoboken,
NJ: Wiley.
INDEX

Figures are indicated by f; tables are indicated by t.

accelerations, 9, 15, 25–27, 61, AISC 360-16; Chapter C, 17


74–75, 78; criteria, 26f; dynamic along-wind response, 9, 29–30, 73
swing, 25–26; frequency- ANSI A58.1, 3
dependent, 26; limits, 27; ANSI/AISC 360-16, 19
low-amplitude, 27; peak, 25–27; Architectural Institute of Japan
static actions, 25; sustained action, (AIJ), 26
25; targets, choice of, 27 AS/NZA 1170.2:2011, 30
acceptance criteria, 13–27; ASCE 7, 14, 17, 24, 36
establishing, 10; building ASCE 7-05, 3
displacements, 20–23; global ASCE 7-10, 3
stability, 13; mean recurrence ASCE 7-16, 2, 3, 14, 20, 29, 47, 49, 62;
intervals, 13–15; nonstructural Appendix C, 15; Appendix CC, 15;
elements, 23–25; occupant C26.5.3, 40; Chapter 27, 52; Seismic
comfort, 25–27; project-specific Design Requirements, 19
performance, 27; stability, 16–19; ASCE 49-12, 47
strength evaluation of the lateral ASCE 7-88, 3
force-resisting system, 19–20 ASCE 7-95, 3
ACI (American Concrete Institute), 3; atmospheric boundary layer (ABL), 9
Committee 375, 70; SP-240-5, 70 auditory disturbances, 27
ACI 301-16, 65 axial effects, 16
ACI 318-14, 17, 19, 65, 68
ACI 363R-10, 65 base movement response, 50f
acoustical disturbances, 2 bending effects, 16
aerodynamic design, 10, 71 Branson equation, 69
aerodynamic modification, 10, building: displacements, 20–23;
71–72, 73f; approximate drift-sensitive, 15; flexure, 21;
effectiveness of, 74t geometry, 10, 78; height, 78;

85
86 INDEX

lean under self-weight, 6; mass, Council on Tall Buildings and Urban


estimating, 31, 32; orientation, 10, Habitat (CTBUH), 47
71–72; porosity, 10; stability, 13;
sway, 27 damping, 2, 53–58; active mass
building codes, 1, 2, 4, 13–14, 19; damper, 58f; aerodynamic, 45–46,
life-safety requirements, 13; 53, 54; aerodynamic, estimates of,
performance objectives, 2, 54; for composite systems, 54;
13–27; regional, 3; unusual compound tank damper, 58f;
structural dynamic properties, 42 direct, 55; for hybrid systems, 54;
building performance: building indirect, 56–58; inherent, 53–54,
periods, 32–33; estimation of, 79; pendulum damper, 57;
30–33; mode shapes, 31, 32–33; piston-style fluid damping device,
preliminary structural analysis, 55f; provided by a structural
31–32; strength checks, 32 system, 53; ratios, 31;
building risk category, 14 recommended critical damping
building services, 1 ratios, 54t; supplemental, 10, 53,
building vertical transportation 54–58; supplemental, strength and
systems, 1, 4 serviceability, 58; tuned liquid
column damper, 56f, tuned mass
cantilever behavior, 18 damper, 57f; tuned sloshing tank
cladding, 23, 24, 61; damage to, 22; damper, 56f; viscous devices, 55;
materials, 24; panel sizes, 23; wall coupling/link beam polymer-
precast concrete, 63 joint damping device, 55; for wind
commercial occupancy; 1-year return induced response, 53–54
period, 26; guidelines, 27 database-enabled design (DED)
computational fluid dynamics (CFD) tools, 30
techniques, 47 Davenport, Alan G., 35–36
computational wind engineering Davenport wind-loading chain,
(CWE), 47, 79 35–36, 71
concrete: cracking, 20, 31–32, 66, deflections, 18, 74; building, 20–23;
68–70; cracking, tensile, 65; first-order, 17; second-order, 17, 19
expected strength, 65–66; deformations, 13, 21; flexural, 21;
force-displacement, 66; heavily- panel zone, 32, 60, 64; permanent,
reinforced, 61; mix, 65; modulus of 1; racking, 22, 24; second-order, 16;
elasticity, 65–66; reduction factors shear, 21, 24, 64; static, 14
for reinforced concrete elements, design load parameters; tabulated, 3
68t; reinforcement, 32; shear, 31, design process, 7–11; comparison of
69–73; stiffness modifiers, 31, results to acceptance criteria, 10;
66–70; stiffness reduction factors, establish performance objectives,
60, 66; strength, 65, 70; 7–8, 13–27; final design, 10–11;
stress-strain curve, 70 overview, 8f; preliminary structural
controlled inelastic response, 19 design, 8; structural modeling and
crosswind: direction, 46; loading, 42; analysis, 10; wind climate
response, 9, 54, 74; vibrations, 30 assessment, 8–9; wind-induced
INDEX 87

loads and responses, 9–10; wind foundation: flexibility, 60, 63;


optimization program, 10 modeling, 63; pinned bases, 63
design validation, 77
diaphragms, 10, 62–63; axial forces, gravity: actions, 19; elements, 31;
63; belt truss levels, 63; flexible, 62; in-plane relative movements, 24;
ground level floors, 63; large load, 17, 19, 31, 62; load-resisting
openings, 54; large reentrant elements, 60
corners, 63; modeling
assumptions, 63; outrigger, 63; height and slenderness ratio, 42
rigid, 62; semirigid, 62–63; holistic optimization, 73–75
setbacks, 63; transfers, 63
differential shortening, 6 irregular plan forms, 42
displacement, 14–15; first-order, 19; ISO 6897, 15, 26, 27
limits, 14; second-order, 19; ISO 10137, 15, 26
vertical, 22f
drift, 31; criteria, 3, 13–15, 23; Krawinkler joint model, 64
horizontal, 24; limits, 25, 27, 32;
nonstructural elements, 23; lateral system, 29, 32; deflections, 1;
performance objective, 31–32; direct load path, 55; linear buckling
story, 18, 20–22, 24; story ratio, analysis, 18; load, 10, 64;
21f, 22 load-resisting, 56, 61–63; mass, 10;
drift damage index (DDI), 24 movement, 16; stiffness, 10, 16, 31;
drift measurement index (DMI), story stiffness, 17
20–23; recommended, 23 linked tower structures, 42
dynamic: effects, 71; mode shapes, live load (LL), 61; code-prescribed, 61;
19; properties, 59; response, 4, in-service, 62
10, 72
main wind force resisting systems
elevators, 4; shaft alignment, 25; shaft (MWFRS), 42, 51
sizes, 23; temporary shutdowns, 25 mass, 61–62; assumptions, 61
ESDU (formerly Engineering Science mean recurrence interval (MRI), 2, 5,
Data Unit), 38 13–15, 39, 46; ∼0.1-year, 25; 1-year,
extreme value analysis, 39–40; annual 15, 25–26, 60; 5-year, 15; 10-year,
maxima, 39; method of 14–15, 23f, 25–26, 60; 25-year, 3,
independent storms, 39; monthly 15, 60; 50-year, 3, 15, 26, 60;
maxima, 39 100-year, 3, 15; 300-year, 3;
400-year, 23f; 700-year, 3;
façade, 24 1700-year, 3; accelerations and
FEMA P-58, 25 motion perception, 15; drift and
final design, 10–11; peer review, 11 displacement, 14–15; DMI criteria
finite-element models, 64 for, recommended, 23f;
floor: diaphragms, 60; mechanical, 62; foundation and lateral system, 14;
plate, 20 large, 24; risk categories, 14f;
88 INDEX

ultimate, 3 member end peer review, 11, 77–78; code, 78;


releases, 61 structural, 78
meteorological data sets, 8 performance-based design (PBD), 79
modeling assumptions, 3 performance objectives, 13–27;
modulus of elasticity (MOE), 65–66; building displacements, 20–23;
concrete, 65–66 life-safety, 19; mean recurrence
modulus of rupture, 68t intervals, 13–15; nonstructural
moment frames, 18 elements, 23–25; occupant
monitoring, 78–79 comfort, 25–27; project-specific
Monte Carlo method, 37 performance, 27; serviceability,
motion predictors, 15 7–8; stability requirements, 7,
16–19; strength evaluation of the
NatHaz Aerodynamic Load Database lateral force-resisting system,
(NALD), 30 19–20; structural strength
natural design analysis, 7 requirements, 7
nonparticipating elements, 60–61 preliminary structural design, 29–33;
nonstructural elements, 1, 23–25; analysis, 31–32; code-based
acceptable drifts for, 23; prescriptive methods, 8; geometric
components and cladding, 24; layout of the structure, 8
damage to, 8; façade, 24; interior project-specific performance, 27
partitions, 24–25; vertical
transportation, 25 quasi-static (background) loads, 45
Notre Dame University, 30
NRC, 30 residential guidelines, 15, 27; 1-year
return period, 26
occupant comfort, 3, 4, 8, 15,
25–27, 41; accelerations, 9, 15, scissor joint model, 66
25–27; acoustical disturbances, second-order effects, 16f
2, 27; improving, 27; visual seismic design, 2, 6, 19, 41
disturbances, 2, 27 serviceability, 3, 5, 8–9, 13, 21, 54;
overturning, 31 accelerations and motion
perception, 15; drift and
P-Delta (second order) effects, 10, displacement, 14–15; -level
16–17, 31–32, 60, 62; amplification, analysis, 60; wind, 14
17; analysis, 17–18; capturing, 16; slender moment column
stability evaluation, 17–18 frames, 17
panel zone deformations, 32, stability, 7, 16–19, 21; acceptance
60, 64; beam-column joints, 60 criteria, 19; coefficient, 18–19;
partitions, 23, 61; audible creaking, evaluation with P-Delta analysis,
25; damage to, 22; detailing, 23; 17–18; global, 18–19, 31; P-Delta
interior, 24–25; materials, 25; visual (second order) effects, 16–17;
cracking, 25 story, 18–19; story stability
pedestrians, 5, 47, 79 coefficient, 17; torsional, 19
INDEX 89

stakeholders, 3–5, 7, 13, 27; architects, structures, 66; strength-level and


4; building occupants, 4–5; façade serviceability analysis, 60
consultant, 4; general public, 5; superimposed dead load (SDL), 61
neighbors, 4; owners, 4; surface wind intensity, 2
pedestrians, 5; structural
engineers, 4; vertical terrain: channeling, 38, 42; far-field
transportation consultant, 4; wind effects, 38; influence of, 38, 73;
engineering consultants, 4 near-field effects, 38; topographic
stiffness, 2; estimating, 31, 59; effects, 38; turbulent wakes, 38
modifiers, 61, 64; modifiers, Torre Shyris, 44f
selecting, 67–70; reduction, 20, 32, twist shape optimization, 73f
60, 66; secant stiffness
approximation, 69f; serviceability ultimate limit states, 46
considerations, 59; tension, 70 ultimate load: combinations, 59, 62;
story drift, 18, 20–22; amount of, 24; conditions, 19
ratio, 21f, 22 USG Corporation, 25
story stability coefficient, 17
strength design, 1, 13, 21, 54; checks, visual disturbances, 2, 27
32; foundation and lateral system, vortex shedding, 30, 42, 46;
14, 19, 32; level analysis, 16, 60 dynamic, 72
Strouhal number, 72
structural analysis: preliminary, wind: along-wind response,
31–32; ultimate, 41 29–30; crosswind response, 30, 42;
structural bay, DMI for a, 22f data models, 9; directionality, 5;
structural modeling and analysis, 10, excitation, 26, 41, 44; nature of,
59–70; building mass, 61–62; 5–6; preliminary estimates, 29–30;
detailed method for selecting return period, 5; “rose,” 5
stiffness modifiers, 67–70; wind climate, 71; analysis, 15, 29–30;
diaphragms, 62–63; expected assessment of, 2, 6, 7, 8–9, 35–40;
strength and modulus of elasticity assessment, site-specific, 9;
of concrete materials, 65–66; assessment, upper-level, 9; data
foundation flexibility, 63; P-Delta sources, 36–37; Davenport
(second order) effects, 62; panel wind-loading chain, 35–36;
zone deformations, 64; primary design criteria, 40; directionality,
lateral load-resisting system and 29, 35; extreme value analysis,
nonparticipating elements, 60–61; 39–40; modeling, 2; storm types,
simplified method for selecting 36–37; strength, 35; terrain, 38
stiffness modifiers, 67; special wind design: nonlinear dynamic
considerations for reinforced analysis, 2; performance
concrete structures, 65–70; objectives, 1–2
stiffness modifiers and behavior of wind-induced loads and responses,
cracked reinforced concrete 9–10, 38; published data on, 9–10
90 INDEX

wind-induced response, 2, 35, 45; data, simulated, 37–38; design


damping levels for, 53–54; evolution, 51; elastic model tests,
resonant component, 44 42; exclusions, 48; high-frequency
wind loads/loading, 4, 21, 29, 41, 61; balance (HFB) approach, 42–46, 48;
code-estimated, 29, 41; criteria, 3; high-frequency pressure
design for, 3; determination of, 2; integration (HFPI), 42, 44, 46, 48;
horizontal, 17; modeling, 2; inclusions, 48; laboratory, 37;
service-level, 31; strength design; loads, 32–33; minimum thresholds,
requirements for, 3, 58; strength- 51–52; multisector method, 49;
level, 58; ultimate, 14, 19 nondirectional, 49; physical, 47;
wind optimization program, 10, predictions, 15; pressure
71–75; aerodynamic design, 71; integration, 45; pressure taps,
aerodynamic modifications, 71; 44–45; procedure, 47; reduced
building geometry, 72–73; scale, 38; reports, 38; required
building orientation, 71–72; input information, 48; rigid
holistic optimization, 73–75; pressure-tapped model, 44; sector
wind speeds, 38; alternative methods, 49; shielding and
minimum design, 9; definition of, influence from surrounding
5; fluctuating component, 5; buildings, 51; static (rigid) model
power spectral density, 5; tests, 42; storm passage, 49;
probability, 3, 8; probability timeline and type,
density function, 5; reference, 3; 47–48; triggers for, 41–42; types of,
standard deviation, 5; ultimate 42–46; typical outputs, 49–51;
wind speed maps, 3 ultimate, 1; up-crossing, 49
wind tunnel tests, 2, 4, 7, 8–10, 29, 31, windstorm types: hurricanes, 37;
33, 41–52, 71; aerodynamic-type synoptic low-pressure systems, 36,
model test, 43–44; benefits of, 41; 39; thunderstorms, 36, 39;
aeroelastic method, 45–46; tornadoes, 36; topographic or
computational wind engineering, thermally driven winds, 36; tropical
38, 47; data and climate, 48–49; cyclones, 36, 37, 39

You might also like