100% found this document useful (1 vote)
3K views

Guidelines For Electrical Transmission Line Structural Loading

Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
100% found this document useful (1 vote)
3K views

Guidelines For Electrical Transmission Line Structural Loading

Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 324

ASCE MANUALS AND REPORTS ON

ENGINEERING PRACTICE NO. 74

Guidelines for
Electrical
Transmission Line
Structural Loading
Fourth Edition

Task Committee on Electrical Transmission Line Structural Loading

Edited by
Frank Agnew, P.E.
ASCE Manuals and Reports on Engineering Practice No. 74

Guidelines for
Electrical Transmission
Line Structural
Loading
Fourth Edition

Edited by Frank Agnew, P.E.

Task Committee on
Electrical Transmission Line Structural Loading

Published by the American Society of Civil Engineers


Library of Congress Cataloging-in-Publication Data

Names: Agnew, Frank, editor. | American Society of Civil Engineers. Task Committee on
Structural Loadings, author.
Title: Guidelines for electrical transmission line structural loading / Task Committee on
Electrical Transmission Line Structural Loading, edited by Frank Agnew, P.E.
Description: Fourth edition. | Reston, Virginia : American Society of Civil Engineers, [2020]
| Includes bibliographical references and index. | Summary: “MOP 74, Fourth Edition,
provides up-to-date design and loading concepts, and applications specific to transmission
line design”-- Provided by publisher.
Identifiers: LCCN 2020018035 | ISBN 9780784415566 (hardcover) | ISBN 9780784483084
(adobe pdf)
Subjects: LCSH: Electric lines--Poles and towers--Design and construction.Load factor design.
Classification: LCC TK3242 .G77 2020 | DDC 621.319/22--dc23
LC record available at https://lccn.loc.gov/2020018035

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 [email protected] 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 https://doi.org/10.1061/9780784415566.

Copyright © 2020 by the American Society of Civil Engineers. All Rights Reserved.

ISBN 978-0-7844-1556-6 (print)


ISBN 978-0-7844-8308-4 (PDF)

Manufactured in the United States of America.

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

Photo credit: Ice photo on cover courtesy of AEP Transmission.


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
useful only occasionally or rarely. It is not in any sense a “standard,” how-
ever, 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
committee 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 man-
uscripts 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 http://www.asce.org/


bookstore.
CONTENTS

PREFACE............................................................................................................ix
ACKNOWLEDGMENTS............................................................................. xiii
1. OVERVIEW OF TRANSMISSION LINE STRUCTURAL
LOADING.................................................................................................... 1
1.0 Introduction.......................................................................................... 1
1.1 Principal Systems of a Transmission Line........................................ 2
1.1.1 Wire System................................................................................. 2
1.1.2 Structural Support System......................................................... 3
1.2 Unique Aspects of Transmission Line Design................................. 4
1.2.1 Tolerance of Failure.................................................................... 4
1.2.2 Designing to Contain Failure.................................................... 5
1.2.3 Coordination of Strengths......................................................... 5
1.2.4 Linear Exposure of Transmission Lines................................... 6
1.3 Load and Resistance Factor Design (LRFD).................................... 6
1.3.1 Reliability-Based Design............................................................ 6
1.3.2 Overview of LRFD...................................................................... 7
1.3.3 Load Factors................................................................................ 8
1.3.4 Strength Factors........................................................................... 8
1.3.5 Sources for Nominal Strengths................................................. 9
1.3.6 Limit States.................................................................................. 9
1.4 Weather-Related Loads..................................................................... 10
1.4.1 Extreme Winds.......................................................................... 10
1.4.2 High-Intensity Winds............................................................... 10
1.4.3 Extreme Ice with Concurrent Wind........................................ 11
1.5 Reliability Concepts for Weather-Related Loads.......................... 11
1.5.1 Mean Recurrence Intervals for Weather-Related Loads...... 11
1.5.2 Relative Reliability and Weather Event MRIs....................... 13
1.5.3 Service Reliability versus Structural Reliability................... 14

v
vi Contents

1.6 Additional Load Considerations..................................................... 14


1.6.1 Construction and Maintenance............................................... 15
1.6.2 Longitudinal and Failure Containment Loads..................... 15
1.6.3 Earthquake Loads..................................................................... 16
1.6.4 Legislated Loads....................................................................... 16
1.6.5 Load Time Signature................................................................ 16
1.7 Wire System........................................................................................ 17
1.8 Examples............................................................................................. 17
1.9 Appendixes......................................................................................... 18
1.10 Draft Prestandard............................................................................ 18
1.11 Incorporation of Changing Data ................................................... 18
2. WEATHER-RELATED LOADS.............................................................. 19
2.0 Introduction........................................................................................ 19
2.1 Wind Loading..................................................................................... 20
2.1.1 Wind Force................................................................................. 20
2.1.2 Air Density Coefficient, Q........................................................ 21
2.1.3 Basic Wind Speed...................................................................... 21
2.1.4 Wind Pressure Exposure Coefficient...................................... 25
2.1.5 Gust Response Factor............................................................... 30
2.1.6 Force Coefficient........................................................................ 34
2.1.7 Topographic Effects.................................................................. 44
2.1.8 Application of Wind Loads to Latticed Towers.................... 48
2.2 High-Intensity Winds........................................................................ 49
2.2.1 Downbursts............................................................................... 49
2.2.2 Tornadoes................................................................................... 51
2.3 Ice and Wind Loading....................................................................... 56
2.3.1 Introduction............................................................................... 56
2.3.2 Categories of Icing.................................................................... 56
2.3.3 Design Assumptions for Ice Loading..................................... 57
2.3.4 Ice Accretion on Wires Due to Freezing Rain....................... 57
2.3.5 Ice Accretion on Structural Members..................................... 65
2.3.6 Unbalanced Ice Loads.............................................................. 66
2.3.7 Ice Accretion on Aerial Marker Balls or Similar Devices.... 66
3. ADDITIONAL LOAD CONSIDERATIONS...................................... 69
3.0 Introduction........................................................................................ 69
3.1 Longitudinal Loads, Line Security, and Failure Containment.... 69
3.1.1 Longitudinal Loads.................................................................. 69
3.1.2 Unbalanced Loads on Intact Systems.................................... 70
3.1.3 Longitudinal Loads due to Non-Intact Wire Systems......... 70
3.1.4 Failure Containment and Line Security Loads..................... 70
3.2 Construction and Maintenance Loads............................................ 71
3.2.1 General....................................................................................... 71
3.2.2 Structure Erection..................................................................... 71
3.2.3 Loads Due to Wire Installation............................................... 73
Contents vii

3.2.4 Maintenance Loads................................................................... 76


3.3 Worker Access and Fall Protection Loads ..................................... 77
3.4 Wind-Induced Structure Vibration.................................................. 77
3.5 Wire Galloping Load Considerations ............................................ 78
3.5.1 Wire Galloping Loads.............................................................. 79
3.5.2 Galloping Mitigation................................................................ 80
3.6 Earthquake Loads.............................................................................. 80
3.6.1 Seismic Hazards........................................................................ 81
3.6.2 Siting and Geotechnical Assessment...................................... 82
3.7 Summary of Additional Load Considerations.............................. 82
4. WIRE SYSTEM.......................................................................................... 85
4.0 Introduction........................................................................................ 85
4.1 Tension Section................................................................................... 86
4.2 Wire Condition................................................................................... 86
4.3 Wire Tension Limits........................................................................... 88
4.4 Calculated Wire Tension................................................................... 89
4.4.1 The Ruling Span Method ........................................................ 89
4.4.2 Structural Analysis of a Single Tension Section ................... 90
4.4.3 Structural Analysis of Multiple Tension Sections ............... 90
4.4.4 Computational Methods.......................................................... 90
4.5 Loads at Wire Attachment Points ................................................... 91
4.5.1 Wire Unit Loads........................................................................ 91
4.5.2 Using Wind and Weight Spans .............................................. 91
4.5.3 Weight Spans on Inclined Spans............................................. 95
4.5.4 Weight Span Change with Blow-Out on Inclined Spans.... 96
4.5.5 Centerline Horizontal Angle versus Wire Horizontal
Angle........................................................................................ 98
5. EXAMPLES................................................................................................ 99
5.0 Latticed Suspension Tower Loads................................................... 99
5.0.1 Design Data............................................................................. 100
5.0.2 Extreme Wind (Chapter 2, Section 2.1)................................ 102
5.0.3 Wind at 30°: Extreme Wind at 30° Yaw Angle
(Chapter 2, Section 2.1)........................................................ 104
5.0.4 Extreme Radial Glaze Ice with Wind (Chapter 2,
Section 2.3)............................................................................. 106
5.0.5 Construction and Maintenance (Chapter 3, Section 3.1)..... 107
5.0.6 Failure Containment (Chapter 3, Section 3.1.4 and
Appendix I, Section 1.3.1).................................................... 109
5.1 Weight Span Change with Blowout on Inclined Spans.............. 110
Shield Wire.........................................................................................111
Conductor......................................................................................... 112
5.2 Traditional Catenary Constant....................................................... 113
Shield Wire........................................................................................ 113
Conductor......................................................................................... 114
viii Contents

A. DEFINITIONS, NOTATIONS, AND SI CONVERSION


FACTORS ................................................................................................ 115
B. RELIABILITY-BASED DESIGN.......................................................... 123
C. AIR DENSITY COEFFICIENT, Q........................................................ 125
D. CONVERSION OF WIND SPEED AVERAGING TIME............... 127
E. SUPPLEMENTAL INFORMATION ON STRUCTURE
VIBRATION............................................................................................ 129
F. EQUATIONS FOR GUST RESPONSE FACTORS.......................... 133
G. SUPPLEMENTAL INFORMATION ON FORCE
COEFFICIENTS...................................................................................... 147
H. SUPPLEMENTAL INFORMATION ON ICE LOADING.............. 167
I. SUPPLEMENTAL INFORMATION REGARDING
LONGITUDINAL LOADS................................................................... 179
J. INVESTIGATION OF TRANSMISSION LINE FAILURES ......... 195
K. HIGH-INTENSITY WINDS................................................................. 209
L. WEATHER-RELATED LOADS FOR ADDITIONAL MRIS.......... 245
M. DRAFT PRE-STANDARD MINIMUM DESIGN LOADS FOR
ELECTRICAL TRANSMISSION LINE FACILITIES...................... 257

REFERENCES................................................................................................. 287
INDEX.............................................................................................................. 301
PREFACE

The American Society of Civil Engineers Task Committee on Electrical


Transmission Line Structural Loading provides design guidance to indus-
try practitioners through the Manuals and Reports on Engineering Prac-
tices. This document, Manual of Practice No. 74, Fourth Edition, is intended
to provide the most relevant and up-to-date information related to trans-
mission line structural loading. It is not intended to be a step-by-step man-
ual or a prescriptive code for direct implementation. Rather, it is intended
to be a resource for development of a loading philosophy for electrical
transmission structures which can be applied to an individual project or at
a regional level. Much of the information contained within this document
can be simplified for particular applications once regional or local climatic
data and reliability levels are determined. The previous editions (1984,
1991, and 2010) have been well received and found wide use as practical
guides to supplement mandatory legal state minimums. Although this
Manual of Practice focuses on applications within the United States, the
concepts presented are applicable worldwide.
In 2012, the ASCE Structural Engineering Institute Committee on Elec-
trical Transmission Structures recognized the need for updates and revi-
sions to Manual of Practice No. 74, Third Edition. The initial intent of the
task committee was to update only sections of the manual affected by
changes to national standards, particularly ASCE 7. As the task committee
commenced review of the impacted sections, they recognized numerous
sections within the manual for which present-day research and recent
industry experience could be applied to significantly improve the content
and organization of the manual. Thus, this resulting fourth edition was
generally rewritten from the third edition.

ix
x Preface

There are several major concept changes in the updated Chapter 1


“Overview of Transmission Line Structural Loading” and Chapter 2
“Weather-Related Loads.” The first of these is the decision to recommend
a 100-year mean recurrence interval (MRI) as the basis for design and pro-
viding the corresponding wind speed and ice thickness maps for the
United States. Additional wind speed maps and combined ice thickness
and wind maps for 50-year and 300-year MRIs are provided in Appendix
L. The additional maps have been included to allow users of this Manual
of Practice to apply wind and ice loads associated with other MRIs as the
previous method for translating loads between MRIs has been discontin-
ued by ASCE 7. Chapter 2 includes some significant changes to compo-
nents of the wind pressure formula, along with an extended discussion of
high-intensity winds, such as downbursts and tornadoes, included in
Appendix K.
Chapter 3 “Additional Load Considerations” and Chapter 4 “Wire Sys-
tem” have been enhanced with additional photos, graphs, and diagrams
to give users of this Manual of Practice a better understanding of the load-
ing concepts and application methodology as presented. Discussions have
also been added to introduce additional loading cases as well as to elabo-
rate on other important transmission structural loading concepts contained
herein.
Chapter 5, “Examples,” has been retained in this edition. The examples
given have been updated to show the methodology of the changes within
other chapters of the document. Chapter 5 has also been expanded in order
to give the user additional guidance on key concepts presented elsewhere
in this document.
Early in the task committee’s work, there was a realization that the elec-
trical transmission line industry would benefit from the development of a
loading standard. As a result, an initial draft of a Transmission Line Struc-
tural Loading Standard document is included in Appendix M of this edi-
tion. This stand-alone draft Pre-Standard is included in this edition in order
for transmission line owners, practitioners, and the public to comment on
the content and form.
The recommendations presented herein reflect the consensus opinion of
the task committee members and are applicable in the context of transmis-
sion line structural loading. Although intended as a guide for lines 69 kV
and greater, the application of the concepts in this document might be justi-
fied at all voltages. The subject matter of this guide has been thoroughly
researched; however, it should be applied only in the context of sound
engineering judgment.
Preface xi

Committee Members
Roberto Behncke Guy Faries Miguel Mendieta
Yair Berenstein Joe Hallman Jacob Merriman
Adam Beyer David Hancock Michael Miller
Ryan Bliss Robert Kluge Garett Muranaka
Gary Bowles Paul Legrand Dave Parrish
Clinton Char Ajay K. Mallik Jeremy Pettus
Ashraf El Damatty Thomas G. Mara Scott Walton
Meihuan Zhu Fulk Julie Matlage C. Jerry Wong
Majid Farahani James McGuire Douglas Zylstra

Respectively submitted,
Task Committee on Structural Loadings
Frank Agnew, Chair
Ron Carrington, Vice Chair
David Boddy, Secretary

ACKNOWLEDGMENTS

The task committee wishes to thank two important groups for their assis-
tance and contributions to this document. The corresponding members of
the ASCE 74 committee provided substantial contributions based on their
expertise in their respective fields. The corresponding members are

Kelly Bledsoe Kathy Jones


Ahmed Hamada Leon Kempner

The second group deserving much praise for their assistance and candid
observations is the Peer Review Committee. It has been a pleasure to work
with these individuals. Their contributions are greatly appreciated.

Ronald Randle, Chair Jean-Pierre Marais


Anthony DiGioia Robert Nickerson
Eric Ho Alain Peyrot
Jon Kell Tim Wachholz
ACKNOWLEDGMENTS

The task committee wishes to thank two important groups for their assis-
tance and contributions to this document. The corresponding members of
the ASCE 74 committee provided substantial contributions based on their
expertise in their respective fields. The corresponding members are

Kelly Bledsoe
Ahmed Hamada
Kathy Jones
Leon Kempner

The second group deserving much praise for their assistance and candid
observations is the Peer Review Committee. It has been a pleasure to work
with these individuals. Their contributions are greatly appreciated.

Ronald Randle, Chair


Anthony DiGioia
Eric Ho
Jon Kell
Jean-Pierre Marais
Robert Nickerson
Alain Peyrot
Tim Wachholz

xiii
CHAPTER 1
OVERVIEW OF TRANSMISSION LINE
STRUCTURAL LOADING

1.0 INTRODUCTION

This Manual of Practice addresses structural loadings to be applied to


transmission lines in the interest of reliable and cost-effective designs in
compliance with regulations, standards, and prescribed design methods.
The following key topics are addressed:
• Uniform procedures and definitions used in the industry for the
calculation of loads. These are intended to facilitate consistency and
communication in the transmission design industry.
• Design procedures that recommend a uniform level of reliability for
transmission lines, as well as a means for increasing or decreasing
this reliability when required. Depending on their importance,
some transmission lines may justify the use of a greater level of
reliability. These procedures may also be used to benchmark the
reliability of existing lines.
• Procedures for calculating design loads and determining their
corresponding load factors. Component and material strengths and
strength factors must also be determined, although the scope of this
manual is limited to general guidance. The designer is directed
toward material-, component-, or product-specific references to
obtain the values to be used with this methodology. Loading
criteria should contain a comprehensive set of loads, as well as
appropriate load factors associated with uncertainty. When prop-
erly coordinated with factored material strengths (which reflect the

1
2 Electrical Transmission Line Structural Loading

variability of materials), the desired levels of reliability can be


obtained. Reliability levels can then be adjusted to meet the perfor-
mance needs and risk tolerance necessary for the facility.
• Failure containment philosophy for structures that are intended to
reduce the probability of cascading failures.
• The most current techniques for quantifying weather-related loads,
namely wind or ice and concurrent wind, that typically control the
design of transmission line structures. High-intensity winds (HIWs)
such as downbursts and tornadoes are also addressed.
• Explanations of wire systems, including how wire tensions and
loads affect a transmission line system.
• Practical examples giving more detail on the application of load
recommendations.
• Appendices containing background information and detailed
discussion of several aspects of transmission line design. This
additional information is intended to supplement the procedures
presented in the main chapters of the manual.
• An appendix containing a draft Pre-Standard for design loads for
electrical transmission line facilities.
Because this Manual of Practice is considered a guideline, it represents
the most current and relevant loading concepts and applications specific
to transmission line design.

1.1 PRINCIPAL SYSTEMS OF A TRANSMISSION LINE

A transmission line consists of two distinct structural systems: the struc-


tural support system and the wire system. The structural support and wire
systems are often considered separately, although they are joined and
respond to loading as a single system. Characteristics and roles of the struc-
tural support and wire systems are described in the following sections.

1.1.1 Wire System


The wire system consists of the conductors and overhead ground/shield
wires and includes all components such as insulators and other hardware
used to attach the wires to the support structures. The majority of the load-
ing on transmission line structures is attributed to the wire system under
gravity and environmental loads (i.e., wind and combined ice and wind).
In addition, much of the unusual behavior and loading challenges faced
by the line designer are generated by the wire system (i.e., unbalanced
tensions, wire vibration, galloping, and broken wire impacts). It is therefore
critical to have a thorough understanding of the wire system to understand
Overview of Transmission Line Structural Loading 3

the loading and response of the overall transmission line system. Loads
applied to support structures are typically described in relation to the
alignment of the transmission line. Longitudinal loads are applied to the
structure in parallel to the transmission line and are caused by unequal
wire tensions on adjacent spans, wire termination, or wire/tower failure.
Transverse loads are applied normally to the transmission line and result
from wind on the wire system (either bare or ice-covered wires) plus the
wire tension resultant because of a line angle (if any). Vertical loads are due
to the self-weight of the wires and attachments, the vertical component of
wire tension, and any accumulated ice.

1.1.2 Structural Support System


The structural support system consists of the towers, poles, guys, and
foundations; and supports the load from the wires, insulators, hardware,
and wire attachments. It also resists wind and combined ice and wind
loads on the wire system, as well as wind and ice loads on the structural
components. The structural support system, or structure, is an essential
element of a transmission line. Each structure is usually evaluated sepa-
rately, although they are joined by the wires, which can transfer load and
cause them to act as a single system. Consideration can be given to evaluat-
ing the structure as part of a system and considering the resisting effects of
the wires that join them.
Structures are typically classified based on the function they perform in
supporting the wire system. The following categories of structures are com-
monly referred to in transmission line design, and are referred to through-
out this manual:
• Tangent structures: Structures that primarily resist the vertical
weight of the wire (and accumulated ice) and transverse loads due
to wind on the structure and wire system. Structures having small
line angles (generally less than 2°) are commonly referred to as
tangents. Although tangent structures can resist substantial trans-
verse and vertical loads at relatively low cost, these structures may
not provide adequate resistance in the case of a high longitudinal
load. Some tangent structure configurations are less resistant to
high longitudinal loads. For these structures an unanticipated
imbalance in longitudinal loading, such as a broken wire, may
initiate a cascade failure.
• Angle or Running Angle structures: Structures marking changes in
line angle along the transmission line where the wires are held in
suspension. Angle structures support the same vertical and trans-
verse loads as tangent structures, as well as significant transverse
loads resulting from wire tensions being applied at an angle.
4 Electrical Transmission Line Structural Loading

• Dead-end structures: Structures designed as termination points for


wires in one or more directions. Although the majority of these
structures under everyday conditions will have wires spanning into
both adjacent spans, they may at some point, usually during
construction and possibly as a result of some type of failure event
in extreme weather conditions, have full design wire tensions
applied on one side (ahead or back span) only. As such, these
structures are often designed to support the full design tension of
all wires pulling longitudinally on one side (ahead or back span) or
face of the structure.
• Strain structures: Structures including both angle strain structures
and in-line strain structures. Structures similar to dead-end struc-
tures, where wire tensions are transferred directly to the structure
but are capable of resisting only unbalanced/differential tensions.
These structures may be used for a variety of reasons, including
supporting tension during wire stringing, accommodation of
clearance limitations, prevention of cascade failure, and resisting
uplift conditions resulting from differences between adjacent spans
(e.g., span length, slope, design tension).

1.2 UNIQUE ASPECTS OF TRANSMISSION LINE DESIGN

1.2.1 Tolerance of Failure


A unique aspect of structural design of electrical transmission line facili-
ties is that failure at some level is acceptable. The acceptable level of risk
of failure often depends on the importance of the transmission line consid-
ered. Transmission grids typically have some level of service redundancy,
which can accommodate failure of a particular transmission line without
any disruption of service. In some cases, a component of a transmission
line can fail and may only damage a small portion of a line, which can be
promptly repaired and service restored with minimal impact to the electri-
cal grid. This is unique in comparison to other engineered structures (e.g.,
buildings, bridges, dams), where a failure could directly result in high
probability of loss of life or substantial property damage. Engineering
judgment should be used to balance reliability of design, minimize the
probability and extent of failure, and provide economical design for the
service life of the transmission line.
There are exceptions where failures resulting in a disruption of service
are to be avoided. Some transmission lines serve critical facilities (e.g., hos-
pitals, emergency services, power plants, cold-start facilities), which may
be in congested, heavily populated areas, or may not have redundancy in
Overview of Transmission Line Structural Loading 5

the grid such as a critical radial feed line. It is recommended to design these
critical transmission lines to a higher level of reliability to reduce the prob-
ability of failure. The designer is directed to Appendix L or ASCE 7-16
(2017), Minimum Design Loads and Associated Criteria for Buildings and Other
Structures, for additional weather loading data for higher MRIs.

1.2.2 Designing to Contain Failure


Recognizing that failure of transmission line components may occur, it
is important to consider failure mitigation or containment in the design of
the line. This failure mitigation or containment is most often addressed by
the use of structures designed for failure containment loads inserted at
regular intervals along the line. This is covered in more detail throughout
the document in the discussion of failure containment. Failure containment
is important to consider in order to prevent the devastating effects of cas-
cading failures, which could result in almost complete destruction of all,
or major portions, of a transmission line.

1.2.3 Coordination of Strengths


The fundamental purpose of coordinating strengths of components or
groupings of components is to limit and contain damage from unantici-
pated loadings in such a way that repairs needed to restore a damaged
transmission line would be faster and more economical. As such, transmis-
sion line systems can be designed in such a way that failure will first occur
in a component which is easy to replace, before other components or the
entire structure, which are more difficult to replace, are damaged. For
example, a tower arm or steel pole arm may be designed to fail at a load
less than that which would cause the entire structure to collapse.
Designing for a sequence of failure through coordination of component
strengths is often difficult as the actual performance may not meet expecta-
tions. This is often due to the large variability in the strength and strength
distributions of the various components of a transmission line. The actual
failure point of any individual component is difficult to predict with accu-
racy, especially when considering the different rates at which materials
deform and deteriorate over time. Many components are also manufac-
tured for nominal strength ratings (e.g., 15,000, 25,000, 36,000, 50,000 lb),
which often results in some components having a greater capacity than
would otherwise be required. Despite the difficulties in achieving coordi-
nation of strengths, it is good design philosophy for the designer to con-
sider how and through what mechanisms their transmission line could
potentially fail, and the consequences resulting from each failure.
6 Electrical Transmission Line Structural Loading

1.2.4 Linear Exposure of Transmission Lines


The fact that transmission lines traverse long distances results in longer
lines generally having a greater possibility of exposure to any low-
probability events. This spatial characteristic of transmission lines results
in longer lines having a lower inherent reliability and an increased prob-
ability of failure compared to shorter ones for the following reasons:
• Uncertainty resulting from exposure to differing terrain, land use,
and natural and artificially manufactured crossings.
• Greater probability of exposure to a weather event striking or
microclimatic condition occurring anywhere along the line. In
addition, the loads produced by weather events are inherently
spatially distributed and can act on transmission lines in singular or
multiple locations with varying intensity over long distances (e.g.,
hurricanes and high-intensity winds).
• Greater probability that a weak component will experience an
extreme loading event, possibly resulting in failure, because there
are more components involved.
By recognizing this spatial characteristic of transmission lines, the
designer must realize that the reliability of a long line is less than that of a
shorter one, all other design parameters being equal.
Having noted the preceding points, it is also evident that it is difficult
to select the appropriate load criteria based on the length of a given trans-
mission line or line section. The result may be that structures and compo-
nents suitable for a specified line length are not suitable for a line of a
different length. One approach is to break long lines into loading zones,
which allows the designer to maintain a consistent reliability level. Another
approach is to establish standards that apply a greater reliability loading
criteria to lines longer than a given length.

1.3 LOAD AND RESISTANCE FACTOR DESIGN (LRFD)

Although this Manual of Practice is a loading manual and not a design


document, the load and resistance factor design (LRFD) concept is pre-
sented to provide context for the recommended loads. The relationship of
load and load factors is presented in relation to various limit states that
may need to be verified in the design or analysis of a transmission line.

1.3.1 Reliability-Based Design


Reliability-based design (RBD) procedures are used to set an acceptable
level of probability of failure and design various components to satisfy this
Overview of Transmission Line Structural Loading 7

target reliability. In terms of statistics, this is achieved by limiting the over-


lap of the lower tail of the probability distribution of the component
strength with the upper tail of the probability distribution of the load or
load effect. However, to carry out a design using RBD procedures, a sig-
nificant amount of material testing and weather data is required to accu-
rately determine the probability distributions of the component strength
and loading. This information is often not readily available. The increased
level of input required, uncertainty in the transfer functions used to calcu-
late loading from weather data, and the resulting loss of accuracy in calcu-
lating the probability of failure of a single component, let alone an entire
system, seldom justify complex RBD procedures. Rather, an attempt is
made to identify and articulate the most useful portions of RBD. The
designer is directed to ASCE Manual of Practice No. 111, Reliability-based
Design of Utility Pole Structures (2006) for additional discussion of RBD. The
practical application of RBD continues to be focused on what can effec-
tively be calculated with statistical weather and strength information in
transmission line design work.
Two key elements of the RBD methodology are (1) methods for adjust-
ing the relative reliability of a line design with respect to ice and/or wind
loads, and (2) techniques for ensuring that structural components have
appropriate strength levels relative to each other. These two elements are
fundamental to maintaining control over the behavior of a line under
extreme loading conditions.

1.3.2 Overview of LRFD


Load and resistance factor design methodology is a simplified approach
to RBD which uses factors to account for uncertainty in the loading and
strength (or other limiting condition). Through this method, a nominal
strength and its corresponding strength reduction factor (φ, sometimes
equal to 1.0) are used to estimate the statistical probability of failure and
determine a nominal design capacity for the component. Conversely, loads
and load effects are increased by load factors (g) that account for the uncer-
tainty and statistical variability of the load, and are used to establish design
loads. Load factors are typically larger than 1, with occasional exceptions.
Each component (i.e., dead load, wind load, construction load) of an
applied load may have a different load factor representing the variability
of that particular load. For example, the load factor for dead loads is typi-
cally much less than the load factor for wind loads due to the relative levels
of uncertainty associated with each type of load. RBD methods are often
used to calibrate appropriate strength reduction factor (φ) and load
factors (γ).
Equation (1-1) shows the basic LRFD concept; in essence, strength must
be greater than, or equal to, load. This relationship should be checked for
8 Electrical Transmission Line Structural Loading

all components of the system with their corresponding strength factors and
limit states, and for all the various load cases and limit states.

·Rn   i Load i (1-1)

where
φ = Strength reduction factor (a different strength factor may be
used for each component and for each limit state);
Rn = Nominal strength of the component;
Load = The appropriate combination of dead loads, wire tensions,
and weather-induced loads, construction and maintenance
(C&M) loads, failure containment, and legislated loads; and
γ = Load factor, which is unique for each load.

1.3.3 Load Factors


Adjustment of the load factor can be applied to the following situations:
• To adjust the reliability with respect to a given weather-related load
of a specific MRI,
• To account for uncertainty or unknowns in the load predictions,
and
• To adjust for the consequences of failure such as worker safety.
For weather-related loads, load factor adjustments should not be com-
bined with MRI adjustments without thorough evaluation. The use of load
factors is appropriate for all non-weather-related loads.

1.3.4 Strength Factors


The purpose of the strength factor, φ in Equation (1-1), is to account for
uncertainty in the coefficient of variation of the strength, COVR, and the
nonuniformity of exclusion limits that currently exist in published formu-
las for nominal strength, Rn. Different strength factors are also used for
different limit states; for example, different factors would be used when
checking the ultimate strength and the cracking limit of a component. The
objective of the line designer is to apply appropriate strength factors to
each component, or groups of components, to control the reliability (or
probability of failure) of the transmission line.
The development or determination of strength factors for components
is beyond the scope of this manual. Guidance on the determination and
use of strength factors is available in IEC 60826, ASCE Manual of Practice
No. 111 (2006), material design standards and guides, and manufacturer
specifications.
Overview of Transmission Line Structural Loading 9

1.3.5 Sources for Nominal Strengths


The following are a selection of standards and guidelines that may be
used to develop nominal strengths (and strength factors) for components
commonly found in transmission lines:
• ASCE 10, Design of Latticed Steel Transmission Structures;
• ASCE 48, Design of Steel Transmission Pole Structures;
• ASCE 104, Recommended Practice for Fiber-Reinforced Polymer Products
for Overhead Utility Line Structures;
• ASCE Manual of Practice No. 123, Prestressed Concrete Transmission
Pole Structures;
• ASCE Manual of Practice No. 141, Wood Pole Structures for Electrical
Transmission Lines;
• ANSI Standards O5.1, O5.2, and O5.3 for wood poles, laminated
timber, wood cross-arms, and braces;
• ANSI C29 Standards for both ceramic and nonceramic insulators;
• ANSI C119 Standards for conductor connectors including dead-end
connections;
• IEEE specifications for various pole line hardware;
• ACI 318, Building Code Requirements for Structural Concrete; and
• AISC 360, Specification for Structural Steel Buildings.

1.3.6 Limit States


While ultimate strength design and other limiting factors are beyond the
scope of this document, it is important to recognize that various limiting
criteria, from deflection to ultimate strength, may control the structural
design of a transmission line. It is therefore important to select the appro-
priate loads for the limit under consideration:
• Failure limit is the point at which a component of the line can no
longer sustain or resist the imposed load. Exceeding this limit will
likely result in the failure of some portion of the line.
• Damage limit is the point at which a component of the line will
suffer permanent damage but may still function, possibly at a
reduced level. Any permanent damage experienced (i.e., plastic
deformation) may affect the future performance and serviceability
of the line. It may also result in a reduced capacity to handle loads
at the damage limit or failure limit levels. Examples of this include
overstressed conductors that may need to be resagged, or hardware
fittings that may be deformed to the point where maintenance is
difficult or impossible. Insulators that have been loaded beyond
their recognized safe working values, and guys which have been
overstressed and require retensioning are also common examples.
10 Electrical Transmission Line Structural Loading

Conventional practice is to ensure that under weather-related


loading, and during construction and maintenance (C&M) opera-
tions, the uses of these components are kept below their damage
limit. This is typically taken as a percentage of their rated (or
nominal) strength. To do this effectively, it is important to under-
stand how the rated or nominal strength is defined by the
manufacturer.
• Deflection or serviceability limits are established to meet a variety of
needs: pole top deflection may be limited for aesthetic purposes;
wire blowout must be checked for various loading conditions to
ensure proper clearances are maintained; changes in conductor sag
and the possible impaired clearances resulting from structure
deflections must be considered; and foundation deflections are
often limited to ensure that excessive nonrecoverable deformations
are not experienced. Due to the nonlinearity of material properties
and the relationship between load and deflection, it is recom-
mended to determine these deflection or serviceability limits using
nonfactored service load cases having a high probability of occur-
rence during the expected life of a transmission line.

1.4 WEATHER-RELATED LOADS

Weather-related loads on transmission line structures and wires are


associated with wind, ice, and temperature or a combination of those loads.
Atmospheric pressure and local topography can influence the characteris-
tics of weather-related loads; these influences should be considered where
appropriate, based on engineering judgment, expert opinion, and past
practice experience.

1.4.1 Extreme Winds


The methodology for the calculation of wind loads on transmission
structures, components, and wires is presented in Chapter 2. The extreme
wind speeds recommended for design are primarily based on the provi-
sions of ASCE 7-16. The general extreme wind loading methodology is
associated with synoptic wind events, including hurricanes.

1.4.2 High-Intensity Winds


Wind loads resulting from convective events, such as thunderstorms,
downbursts, and tornadoes, differ in magnitude, load distribution, and
duration from the extreme winds that transmission structures are normally
designed to resist. Damage and failure resulting from HIWs have been
frequently described for many locations around the world. Some locations
Overview of Transmission Line Structural Loading 11

experience more of these events than others, and engineering judgment


and past experience with the area should be considered when applying
HIW methods. A general description of HIW is given in Chapter 2, while
approaches to design for HIW (including downbursts and tornadoes) are
given in Appendix K.

1.4.3 Extreme Ice with Concurrent Wind


Extreme ice with concurrent wind speeds are presented in Chapter 2.
The formation of ice on transmission structures is generally referred to as
ice accretion and is often more important for the wires than for the struc-
tures. Ice accretion on a transmission line often imposes substantial vertical
loads on the structural system. The resultant load on the wires causes sig-
nificantly greater wire tensions compared to bare wire conditions. Chap-
ter 2 provides a basis for estimating the thickness of ice on conductors and
shield wires, as well as the concurrent wind speeds and air temperatures
to be considered with ice accretion.

1.5 RELIABILITY CONCEPTS FOR WEATHER-RELATED LOADS

1.5.1 Mean Recurrence Intervals for Weather-Related Loads


It is customary to associate extreme values of weather events with some
MRI, which is commonly referred to as a return period. For example, a MRI
of 100 years is associated with an event which, on average, has a probabil-
ity of 1/100 (or 1%) of being met or exceeded in any given year or a 39%
chance of being met or exceeded at least once during a 50-year period.
However, because extreme weather events are not necessarily evenly
spaced over time, some 50-year periods may pass with no weather equal
to or exceeding the level associated with the 100-year MRI. Conversely,
some 50-year periods may experience multiple events equal to or exceed-
ing the level associated with the 100-year MRI. The MRI analysis assumes
that the probability of the event occurring does not vary over time and is
independent of past events. The following formula can be used to calculate
the probability of exceeding the level corresponding to a given MRI over
an N-year period.


( ) 
N
1
Probability of Exceedance in N years = 1 − 1 − (1-2)
 MRI

Table 1-1 includes a column showing the probability of exceedance of


various MRIs one or more times during a span of 50 years. For example,
the probability of exceeding the wind speed associated with a 50-year MRI
Table 1-1.  Exceedance Probability for Various MRIs.

Location of wind and ice


Probability the Probability the MRI Probability the MRI
maps shown in ASCE
MRI load is load is exceeded at load is exceeded at
MOP 74
MRI exceeded in any least once in 50 years least once in 100
Typical conditions (years) one year (%) (%) years (%) Wind Ice
e
Temporary or emergency 10 10 99 99+ —
restoration, service checksa
e
Temporary or emergency 25 4 87 98 —
restoration, service checksa
Historically used MRIc 50 2 64 87 Appendix L Appendix L
Recommended MRI 100 1 39 63 Chapter 2 Chapter 2
Enhanced reliability 200 0.50 22 39 — —
Enhanced reliability 300 0.30 15 28 Appendix L Appendix L
Enhanced reliability 400 0.25 12 22 — —
d
Enhanced reliability 500 0.20 10 18 —
b
Enhanced reliability 700 0.14 7 13 —
a
Service checks include clearance checks, blowout checks, insulator uplift, and deflections.
b
ASCE 7-16, Chapter 26.
c
Previous editions of this manual provided wind speeds associated with a 50-year MRI for a majority of the continental United States,
with MRIs of the hurricane-prone regions in the range of 50 to 90 years. See Section 1.5.1 and Chapter 2 for further explanation.
d
ASCE 7-16, Chapter 10.
e
ASCE 7-16, Appendix C Commentary.
Overview of Transmission Line Structural Loading 13

at least once in a 50-year period is [1– (1 – (1/50))50] = 0.64, or 64%. It should


also be noted that there is an almost 1 in 4 probability of a 200-year MRI
event occurring at least once in 50 years, [1– (1 – (1/200))50] = 0.22, or 22%.
Maps of extreme wind and extreme ice with concurrent wind are pre-
sented in Chapter 2. The extreme wind and ice loads are provided for
specific MRIs. Applying loads for a different MRI can be used to vary the
target reliability of the design. For weather-related loads, the recommended
method to adjust the relative reliability level is to use a load corresponding
to the MRI of interest. This is most accurately done by using the extreme
wind and extreme ice with concurrent wind maps provided, as the rela-
tionship between extreme weather-related loads and MRI varies spatially
throughout the United States. The use of relative reliability load factors to
transform basic wind speeds and ice thicknesses to different MRIs is not
promoted in this edition. Rather, maps relating to specific MRIs have been
proposed to improve consistency of risk, as detailed in Chapter 2.
This edition of the manual recommends using 100-year MRI extreme
wind and ice loads as the basis for design. Extreme wind and ice maps for
additional MRIs are provided in Appendix L and ASCE 7-16.
The previous edition of this manual presented a basic wind speed map,
taken from ASCE 7-05 (2005), which was commonly interpreted as a
50-year return period wind speed map. However, the return period varied
in hurricane-prone regions from 50 to 90 years. Further information is pro-
vided in the commentaries to the wind load chapters in ASCE 7-05 and
ASCE 7-10 (2010).
The wind speed maps in this edition of this manual are adopted from
ASCE 7-16. The maps in ASCE 7-16 address the issue of risk consistency
between nonhurricane and hurricane regions and incorporate additional
wind data obtained since the development of the previous map. The
100-year MRI wind speed map in Chapter 2 provides the same risk consis-
tency and reliability across the United States.

1.5.2 Relative Reliability and Weather Event MRIs


The reliability of a transmission line, a single structure, or a single com-
ponent can be adjusted up or down by choosing higher or lower MRIs,
respectively. Several factors should be considered prior to adjusting the
design MRI for a transmission line or one of its components. For example,
considering the critical importance of the facilities served, or the generation
source, the line owner may desire an increased level of reliability and there-
fore select design loads associated with a higher MRI. Unique structures
that may be difficult to replace, structures at major road or river crossings,
or structures in a high-density urban area may also justify an increased
level of reliability.
14 Electrical Transmission Line Structural Loading

Conversely, for temporary lines, emergency lines, existing transmission


lines with relatively short life spans, or where the consequences of failure
are deemed relatively inconsequential, a decreased level of reliability may
be deemed appropriate. The applicability and requirements of legislated
loads must be considered when reducing the MRI as to not violate any
minimum requirements. Sound engineering judgment should always be
used when deviating from the 100-year MRI recommended in this Manual
of Practice.
The relative probability of exceedance of a load is inversely proportional
to the design load MRI (Peyrot and Dagher 1984). Hence, doubling the MRI
of the design load reduces the probability of exceedance in any one year
by a factor of 2, as shown in Table 1-1.
The concept of relative reliability can be used as a tool to approximately
adjust design reliability as it is currently very difficult to accurately calcu-
late the probability of failure of a line. Powerful mathematical tools which
can predict transmission line reliability are available, but such detail is
often not justified considering the linear extent of a line, the redundancy
normally included in transmission systems, and the availability of the data
required to perform such an analysis. Specifically, predicted probabilities
of failure are in error if the uncertainties in the probability distribution of
climatic events, structure strengths, component strength, and transfer func-
tions converting climatic events into loads are not adequately considered.
Until more information becomes available to resolve the load and strength
data issues, it is recommended that the concept of relative reliability be
used, as an admittedly approximate tool, to adjust structural reliability.

1.5.3 Service Reliability versus Structural Reliability


There are many events that affect the service reliability of a transmission
line. Some are related to electrical events and may be controlled by station
equipment rather than transmission line design. Others may be
storm-related, such as electrical flashovers experienced during lightning
events. These aspects are not included in the scope of this manual but
contribute to the overall electric service reliability of the transmission line.
When determining the desired service reliability, one must separate the
electrical service reliability from the structural reliability.

1.6 ADDITIONAL LOAD CONSIDERATIONS

There are several other types of loads, loading conditions, and informa-
tion on the time signature of loading (see Section 1.6.5) that may need to
be considered in the design or analysis of a transmission line and its com-
ponents. Chapter 3 includes a detailed discussion of some of these other
Overview of Transmission Line Structural Loading 15

types of loading, including construction and maintenance loads, and lon-


gitudinal and failure containment loads. Chapter 3 also includes discus-
sions regarding load effects due to galloping wires, as well as potential
problems due to the effects of vibration of structural members and seismic
events.

1.6.1 Construction and Maintenance


Construction loads are loads the structure must resist during assembly,
erection, and throughout the installation of shield wires, insulators, con-
ductors, and line hardware.
Maintenance loads are experienced by the structure as a result of inspec-
tion, replacement of structure components, or alteration of the supported
facilities. Structure maintenance loads consider the effects of workers on
the structure, as well as load effects on adjacent structures due to tempo-
rary modifications such as guying to repair or replace a structure. Person-
nel safety should be the paramount factor when establishing all loads,
especially C&M loads. Construction and maintenance loads and conditions
are discussed in more detail in Chapter 3.

1.6.2 Longitudinal and Failure Containment Loads


Structures may experience longitudinal loads (as a result of tension dif-
ferences in adjacent spans) in a variety of scenarios. These loads should be
included in the design of structures when warranted. These scenarios
include, but are not limited to
• Inequalities of wind and/or ice on adjacent spans (e.g., ice on the
back span with no ice on the ahead span),
• Ice- and/or wind-load-induced longitudinal imbalance on consecu-
tive spans with significant difference in length (e.g., tangent span
over a valley),
• Wire breakage,
• Construction or maintenance loading scenarios (e.g., wire caught in
the stringing block),
• Hardware swing constraints that restrict balancing of tensions,
• Insulator failure, and
• Structural and hardware component failure.
Longitudinal loading events can create severe load imbalances in the
wire system capable of causing the partial or complete failure of the adja-
cent supports. Propagation of the failure to a multitude of support struc-
tures along a transmission line as a result of longitudinal load imbalances
is generally referred to as a cascade failure. These cascade failures cause
significant damage and high economic losses as complete sections of a
16 Electrical Transmission Line Structural Loading

transmission line may be destroyed, resulting in weeks or months of repair


(Ostendorp 1997). Cascade-type failure events can be prevented or mini-
mized by the inclusion of certain longitudinal load cases and scenarios.
Longitudinal loads and approaches to failure containment are discussed
in further detail in Chapter 3.

1.6.3 Earthquake Loads


Historically, transmission line structures have not failed due to inertial
loads caused by earthquakes. Decades of industry experience and records
support the observation that little, if any, damage is observed on transmis-
sion structures after an earthquake event. Therefore, the structural design
of transmission line structures need not consider inertial loads associated
with earthquakes.
While the design of transmission line structures will not be controlled by
earthquake-induced inertial loads, structural damage could occur as a result
of secondary effects, such as damaged foundations or tiebacks, use of rigid
post-type insulators, minimal wire slack between structures, soil liquefaction
and landslides. When locating or designing a new transmission line route,
recognition of various seismic hazards should be considered to minimize
potential damage. Additional descriptions of hazards and potential second-
ary effects due to ground motion are provided in Chapter 3.

1.6.4 Legislated Loads


Legislated loads refer to minimum loads specified by federal, state, or
municipal codes and legislative or administrative acts. In the United States,
the most common source of legislated loads is the National Electrical Safety
Code (NESC) (IEEE 2017), which specifies minimum loading requirements
for safety as specified in Rule 010.
Legislated codes may require load cases in addition to those imposed
by atmospheric conditions, such as those caused by anticipated construc-
tion and maintenance activity. In addition to loads, allowable material
strengths, load factors, and calculation methodologies may also be speci-
fied. Legislated loads are not included in the scope of this manual.

1.6.5 Load Time Signature


When prescribing a load for a given situation, information in addition
to the magnitude is required to achieve the desired transmission line per-
formance. The designer may also need to consider the time history of the
load, or load signature, as the resistance of materials can vary with load
Overview of Transmission Line Structural Loading 17

duration, dynamic load characteristics, and load intervals (impact or


cyclical).
The following examples illustrate the response of different materials to
various loading characteristics:
• Wood, nonceramic insulator, and fiber-reinforced composite
member strengths must be reduced when exposed to long-duration
loads, such as those that might occur during an icing event or
recurring cumulative load events;
• Metallic materials exhibit fatigue when exposed to cyclical loading
events such as vibration;
• Impact loads, such as those which may occur during a conductor
galloping event, can be detrimental to ceramic materials; and
• Hardware and connections can be loosened by vibration.
As can be seen from this list, additional checks may be required depend-
ing on the load signature and affected component materials.

1.7 WIRE SYSTEM

To determine the loads on transmission structures accurately, it is neces-


sary to understand how tensions are generated in the wire systems, and
how the resulting loads are imposed on the support systems. These aspects
are described and discussed in Chapter 4. Wire tension changes with tem-
perature, time (creep), ice, wind, the flexibility of the support, and C&M
operations. Chapter 4 discusses the manner in which the wire system
responds to these unit wire loads and some of the assumptions that may
be used for determining the loads at the structure attachment points. Chap-
ter 4 also provides information regarding
• Effects of creep and heavy loads on wire tensions,
• The need to keep wire tensions within certain limits, and
• Simplifications and assumptions that may be used for determining
wire tensions and resulting loads at structure attachments.

1.8 EXAMPLES

The application of the recommended loadings in Chapters 2 and 3 is


illustrated for a typical suspension tower in Chapter 5. Appendix F con-
tains detailed example calculations for the gust response factors for the
same tower.
18 Electrical Transmission Line Structural Loading

1.9 APPENDIXES

A number of appendixes are included, which provide additional infor-


mation to supplement the various loading topics covered in the main body
of the manual. Where applicable, the relevant appendix is referenced in the
corresponding chapter of the manual. Additional information, background
and history, and derivations are available for the majority of the recom-
mended loading provisions.

1.10 DRAFT PRESTANDARD

One of the goals of this edition of the manual was to create a draft pre-
sentation for a standard for Transmission Line Loading. The ultimate goal
is to develop a standard for Design Loads for Electrical Transmission Line
Facilities. The initial draft of this prestandard document is included in
Appendix M in order to disseminate it for public review and comment, and
to gather and consider industry consensus for future revision and
publication.

1.11 INCORPORATION OF CHANGING DATA

When applying this Manual of Practice to design, the line designer must
recognize that the contents of this manual were developed based on his-
torical performance of both environmental loading as derived from
recorded meteorological data and of the performance of transmission line
structures in response to this loading. In other words, the knowledge accu-
mulated and implemented in this and other manuals and guidelines is
based on data and performance of the past. The database of weather
records is continuously expanding with each passing year and as more
weather stations are installed. The importance of obtaining new meteoro-
logical and performance data is more important with global climate
change, as the effects of these changes on the electrical system are largely
unknown. Proactive efforts to gather and apply additional data should be
engaged where possible and practical to ensure electrical service continuity
relative to structural loading.
CHAPTER 2
WEATHER-RELATED LOADS

2.0 INTRODUCTION

This chapter discusses weather-related loads on transmission line struc-


tures and wires. These loads are associated with wind or a combination of
ice and wind, referred to as ice with concurrent wind. Temperature, atmo-
spheric pressure, and local topography influence the magnitude of
weather-related loads. These influences should be considered where
appropriate.
A standard wind pressure formula applicable to transmission lines is
presented. The formulation of design wind pressures recommended in this
chapter is primarily based on the provisions of ASCE 7-16 (2017). The wind
equations presented in this Manual of Practice are developed from infor-
mation currently available in the engineering community and provide
practical methods for design resulting in adequate levels of performance.
However, due to several challenges, as well as the spatial extent of trans-
mission lines, precise wind load prediction is difficult. Some of these chal-
lenges include the uncertainty in quality and quantity of wind data, the
transfer functions to convert wind velocity to wind pressure, and the vari-
ous meteorological mechanisms generating extreme wind (e.g., large-scale
pressure systems, convective storms). It has been well documented that the
load effects resulting from convective events, such as thunderstorms,
downbursts, and tornadoes, differ from those for which transmission struc-
tures are normally designed. These types of winds are often referred to as
high intensity winds (HIWs).
Ice with concurrent wind loads are also described. The formation of ice
on transmission line components is generally referred to as ice accretion.
This manual provides regional values of ice thickness due to freezing rain

19
20 Electrical Transmission Line Structural Loading

and the wind speeds and air temperatures to be considered in combination


with ice accretion.
Supplemental information on converting wind speed to pressure, wind
speed average times, gust response factors, force coefficients, and ice accre-
tion is given in Appendixes C, D, F, G, and H, respectively. Appendix K
provides advice and methods regarding the consideration of HIWs in the
design of transmission structures. Appendix L provides weather-related
loads for additional MRIs which may be of interest in design.

2.1 WIND LOADING

2.1.1 Wind Force


The wind force acting on the surface of transmission line components can
be determined by using the wind force expression shown in Equation (2-1)

F = QKzKzt(V100)2GCfA (2-1a)

or

F = QKzKzt(VMRI)2GCfA (2-1b)

where
F = Wind force in the direction of wind unless otherwise speci-
fied [lb (N)];
G = Gust response factor for conductors, ground wires, and
structures as specified in Section 2.1.5;
Cf = Force coefficient values as recommended in Section 2.1.6;
A = Area projected on the plane normal to the wind direction [ft2
(m2)];
Q = Air density coefficient defined in Section 2.1.2;
Kz = Wind pressure exposure coefficient, which modifies the
reference wind pressure for various heights above ground
based on different exposure categories defined in Section
2.1.4. The values are obtained from Equation (2-3) or
Table 2-2;
Kzt = Topographic factor obtained from Equation (2-16) in Section
2.1.7;
V100 = Reference 3-second gust wind speed for 100-year MRI [mph
(m/s)], obtained from Figure 2-1 in Section 2.1.3; and
VMRI = Reference 3-second gust wind speed for selected MRIs [mph
(m/s)], obtained from ASCE 7–16 or Appendix L of this
manual.
Weather-Related Loads 21

The wind force calculated from Equation (2-1) is based on the selection
of appropriate values of wind speed, wind pressure exposure coefficient,
gust response factor, and force coefficient. These parameters are discussed
in subsequent sections. Wire tension corresponding to the wind loading
should be calculated using the wire temperature that is most likely to occur
at the time of the extreme wind loading event.

2.1.2 Air Density Coefficient, Q


The air density coefficient, Q, converts the kinetic energy of moving air
into potential energy of pressure. For wind speed [mph (m/s)] and pres-
sure [psf (Pa)], the recommended value is

Q = 0.00256 mph to psf (0.613 m/s to Pa) (2-2)

The nominal value of Q in Equation (2-2) reflects the specific weight of


air for a standard atmosphere [i.e., temperature of 59 °F (15 °C) and sea
level pressure of 14.7 psi (101.325 kPa)]. For some cases, such as line
upgrading/re-rating in high elevations, the effects of temperature and
elevation (atmospheric pressure) on the value of Q may be considered.
Sufficient meteorological data should be available to justify a different
value of the air density coefficient for a specific design application. Varia-
tion of Q with temperature and elevation is discussed in Appendix C.

2.1.3 Basic Wind Speed


In the United States, the basic wind speed is the 3-second gust wind
speed at 33 ft (10 m) above ground in open country terrain (Exposure Cat-
egory C, as defined in Section 2.1.4.1). ASCE 7-16 provides basic wind
speeds associated with various MRI in the form of contour maps. Wind
speed maps for MRI of 10, 25, 50, 100, 300, 700, 1,700, and 3,000 years are
provided in ASCE 7-16.
In the previous versions of this manual (ASCE 1991, 2010a), a single map
of basic wind speeds was provided for the United States. Relative reliabil-
ity factors were then provided to transform the basic wind speed, which
was associated with a nominal 50-year MRI, to additional MRIs of interest
(i.e., 25 years, 200 years). This process assumed that the relationship of
wind speed and MRI is consistent among all regions. The study of wind
climatology through longer historical records, as well as improvements in
data collection and processing, has led to a different approach in the wind
engineering community.
Based on an updated analysis of historical wind data in the United States,
wind speed maps corresponding to several MRIs have been developed and
are presented in ASCE 7-16 (e.g., Pintar et al. 2015). The wind speed maps
22 Electrical Transmission Line Structural Loading

included in the main body of ASCE 7-16 represent MRI associated with
factored strength design (i.e., 300, 700, 1,700, 3,000 years) based on target
reliability levels for building occupancy and function. Wind speed maps not
associated with ASCE 7 strength factors for building occupancy and func-
tion (i.e., 50, 100 years) are available in Appendix C of ASCE 7-16.
The wind speed map recommended by this manual corresponds to a
MRI of 100 years, consistent with the 100-year MRI wind speed map in
ASCE 7-16, and is shown in Figure 2-1. This map provides wind speeds
that are consistent with the intent of the basic wind speed map in the previ-
ous editions of this manual with respect to reliability. Values are nominal
design 3-second gust wind speeds in miles per hour (m/s) at 33 ft (10 m)
above ground in Exposure Category C terrain. Linear interpolation
between contours is permitted. It is acceptable to use the last wind speed
contour of the coastal area for islands and coastal areas beyond the last
contour. Further examination is recommended for unusual wind condi-
tions in areas identified as Special Wind Regions. Additional wind speed
maps for MRIs of 50 years and 300 years are provided in Appendix L.
Selection of a MRI of a value other than 100 years may be desirable for
certain applications.
The entire state of Hawaii is defined as a Special Wind Region on the
current wind speed maps. This is due to the extreme topographic condi-
tions found throughout the Hawaiian Islands. Significant research on wind
speedup due to topographic features has been carried out for the Hawaiian
Islands (e.g., Chock and Cochran 2005), resulting in maps of the topo-
graphic factor Kzt. These detailed maps are publicly available through the
Department of Accounting and General Services for the state of Hawaii.
Following a review of the Kzt maps for Hawaii, the following wind speeds
are recommended for a MRI of 100 years: (1) A wind speed of 105 mph
(47  m/s) for regions indicated as Kzt ≤ 1.5, and (2), a wind speed of
( ) ( )
86 mph K zt or 38 m / s K zt for regions indicated as Kzt > 1.5.
In certain regions of the country, such as mountainous terrain, topo-
graphical characteristics may cause significant variations in wind speed
over short distances. See Section 2.1.7 for further discussions on topo-
graphic effects. These variations in wind speeds are dependent upon local
effects and cannot be effectively shown on the maps. In addition, Special
Wind Regions indicate that wind speeds in these regions may vary signifi-
cantly from those values indicated on the map. In these cases, the designer
should consult a meteorologist or wind engineer to establish a design
wind speed.
Weather-Related Loads 23

Figure 2-1. 100-year MRI 3-second gust wind speed map [mph (m/s)] at 33 ft
(10 m) aboveground in Exposure Category C (Continued)
Source: ASCE (2017).
24 Electrical Transmission Line Structural Loading

Figure 2-1. (Continued) 100-year MRI 3-second gust wind speed map [mph
(m/s)] at 33 ft (10 m) aboveground in Exposure Category C
Source: ASCE (2017).

2.1.3.1 Use of Local Wind Data  It is possible to determine the basic wind
speed using regional wind data for a specific location. ASCE 7-16 provides
criteria for the use of regional meteorological data.

2.1.3.2 Wind Speed Conversion  A conversion procedure to obtain wind


speeds of different averaging times is described in Appendix D.
Weather-Related Loads 25

2.1.4 Wind Pressure Exposure Coefficient


It is recognized that wind speed varies with height due to interaction
(friction) with the surface of the earth. This is referred to as the atmospheric
boundary layer and is dependent on ground surface roughness as charac-
terized by the various exposure categories described in the next section.
The wind pressure exposure coefficient, Kz, used in Equation (2-1) and
defined in Equation (2-3) (Section 2.1.4.2), modifies the basic wind pres-
sure, taking into account the effect of height and terrain.

2.1.4.1 Exposure Categories  The following terrain roughness or exposure


categories are recommended for use with this manual and are specified in
ASCE 7-16. The recommended exposure category for transmission lines is
Exposure C, unless the criteria of Exposures B or D can be met and there is
a reasonable expectation that the exposure category will not change over
the life of the transmission line. The entire state of Hawaii has detailed
Exposure Category maps to correspond to the topographic maps described
in Section 2.1. Use of these maps is recommended where applicable.

Exposure B  This exposure is classified as urban and suburban terrain,


densely wooded areas, or terrain with numerous, closely spaced obstruc-
tions having the size of single-family dwellings or larger. A typical view of
terrain representative of Exposure B is shown in Figure 2-2. Use of Expo-
sure B shall be limited to wind directions for which representative terrain

Figure 2-2. Typical terrain representative of Exposure B.


26 Electrical Transmission Line Structural Loading

extends either 2,600 ft (792 m) or 20 times the height of the transmission


structure, whichever is greater.
In the use of Exposure B, the question arises as to what is the longest
distance of flat, unobstructed terrain located in the middle of a suburban
area permitted before Exposure C must be used. A guideline is 600 ft (180
m) or 10 times the height of the transmission structure, whichever is less,
as the length of intermediate, flat, open country allowed for continued use
of the Exposure Category B. The surface conditions required for the use of
the Exposure Category B are illustrated in Figure 2-3.

Figure 2-3. Surface conditions required for the use of Exposure Category B.

Exposure C  This exposure is defined as open terrain with scattered


obstructions having heights less than 30 ft (9.1 m). This category includes
flat, open country, farms, and grasslands. A typical view of terrain repre-
sentative of Exposure C is shown in Figure 2-4. This exposure category
should be used whenever terrain does not fit the descriptions of the other
exposure categories. It should also be noted that this exposure is represen-
tative of airport terrain, where most wind speed measurements are
recorded.

Exposure D  This exposure is described as flat, unobstructed areas directly


exposed to wind flowing over open water for a distance of at least 5000 ft
(1524 m) or 20 times the height of the transmission structure, whichever is
greater. Note that the ASCE 7-16 wind speed map requires the use of Expo-
sure D along the hurricane coastline (Vickery et al. 2010). A typical view of
terrain representative of Exposure D is shown in Figure 2-5. Shorelines for
which Exposure D also applies include inland waterways, the Great Lakes,
and coastal areas of California, Oregon, Washington, and Alaska. The
Exposure Category D applies only to structures directly exposed to bodies
of water and coastal beaches. The surface conditions required for the use
of Exposure Category D are illustrated in Figure 2-6.
Weather-Related Loads 27

Figure 2-4. Typical terrain representative of Exposure C.

Figure 2-5. Typical terrain representative of Exposure D.


28 Electrical Transmission Line Structural Loading

Figure 2-6. Surface conditions required for the use of Exposure Category D.

2.1.4.2 Exposure Coefficient Equations  Values of the wind pressure


exposure coefficient, Kz, are calculated using Equation (2-3).
2
 z a
Kz = 2.01  h  for 33 ft ≤ zh ≤ zg (2-3)
 z g 
where
α = Power law coefficient for gust wind (see Table 2-1);
zh = Effective height; and
zg = Gradient height (see Table 2-1).

The effective height, zh, is discussed in Section 2.1.4.3. The gradient


height, zg, defines the thickness of the atmospheric boundary layer. Above
this elevation, the wind speed is assumed to be constant. The power law
exponent, α, accounts for the shape of the wind speed profile with respect
to height. Values for the power law exponent and corresponding gradient
heights are listed in Table 2-1 for each exposure category. Table 2-2 can be
used to determine Kz for heights up to 200 ft (60 m) above ground.
Weather-Related Loads 29

Table 2-1.  Power Law Exponent for Gust Wind Speed and
Corresponding Gradient Height

Exposure Category α zg (ft)

B 7.0 1200

C 9.5 900

D 11.5 700

Table 2-2.  Wind Pressure Exposure Coefficient, Kz.

Wind Pressure Exposure Coefficient, Kz


Effective Height,
zh (ft) Exposure B Exposure C Exposure D

0−33 0.72 1.00 1.18

40 0.76 1.04 1.22

50 0.81 1.09 1.27

60 0.85 1.14 1.31

70 0.89 1.17 1.35

80 0.93 1.21 1.38

90 0.96 1.24 1.41

100 0.99 1.27 1.43

120 1.04 1.32 1.48

140 1.09 1.36 1.52

160 1.13 1.40 1.55

180 1.17 1.43 1.59

200 1.20 1.46 1.62


30 Electrical Transmission Line Structural Loading

The effects of the wind pressure exposure coefficient on wind force for
the different exposure categories are significant. It is essential that the
appropriate exposure category be selected after careful review of the sur-
rounding terrain. It is recommended that Exposure C be used unless the
designer has absolutely determined that Exposure B or Exposure D is more
appropriate. The transfer of the basic wind speed between exposure cate-
gories should only be used with sound engineering judgment.

2.1.4.3 Effective Height  The effective height, zh, is theoretically the height
from ground level to the center of pressure of the wind load. The effective
height is used for selection of a wind pressure exposure coefficient, Kz
[based on Equation (2-3) or Table 2-2], and calculation of the gust response
factors, Gw or Gt [Equations (2-4) and (2-5) in Section 2.1.5].
The effective height of a conductor and ground wire subjected to wind
and ice with concurrent wind, is influenced by the blow-out swing of the
wires and insulators. However, for structural design purposes, the effective
heights of all the wires can be approximated as the average height above
ground of all the wire attachment points to the structure.
The wind pressure exposure coefficient varies over the height of the
structure. Transmission structures may be divided into sections, where the
effective height, zh, is the height to the center of each section. For some
structures, a second or simpler alternative for structure heights 200 ft (60
m) or less is to assume one section and use two-thirds of the total structure
height as the effective height. This alternative will apply a uniform wind
pressure over the height of the structure. Sound engineering judgment
should be used in the application of this approach.

2.1.5 Gust Response Factor


The gust response factor accounts for the dynamic effects and the cor-
relation of gusts on the wind response of transmission line components. It
is recognized that gusts generally do not envelop the entire span of wires
between transmission structures, and that some reduction reflecting the
spatial extent of gusts should be included in the calculation of the wind
load. Both the dynamic effects and correlation have been incorporated in
the gust response equations developed by Davenport (1979).
It should be noted that the gust response factor is different from the gust
factor, which is used by some electric utilities in their wind loading criteria.
The gust factor is the ratio of the gust wind speed at a specified short dura-
tion (e.g., 3 seconds) to a mean wind speed measured over a specified
averaging time (e.g., hourly mean or 10- minute mean). The gust response
factor is the ratio of the peak load effect on the structure or wires to the
mean load effect corresponding to the design wind speed. Therefore, the
gust factor is a multiplier of the mean wind speed to obtain the gust wind
Weather-Related Loads 31

speed, whereas the gust response factor is a multiplier of the design wind
load to obtain the peak load effect. The gust response factors described here
replace the use of traditional gust factors.
The equations for gust response factors (Davenport 1979), described in
Appendix F, were originally developed based on an hourly mean wind
speed, as discussed by Behncke and Ho (2009). In the previous edition of
this manual, the original Davenport gust response factors were modified
to be consistent with the definition of the basic wind speed in ASCE 7,
which is the 3-second gust at 33 ft in open country terrain (Exposure Cat-
egory C). However, the equations retained parameters whose definitions
have been improved upon since their original development. Thus, the
equations for the gust response factors have been modified for this edition
of the manual. Most notable are the removal of the parameters κ (surface
drag coefficient) and E (exposure factor); these have been replaced with
modern parameters used in the description of atmospheric boundary layer
wind. The modified approach involves the calculation of the turbulence
intensity of the wind at the effective height of the structure or wires, as well
as the use of separate peak factors for the background and resonant com-
ponents of the dynamic response. The revisions to the methodology reflect
the state of the art in the calculation of wind loads on structures and are
consistent with the methodology for atmospheric boundary layer winds
applied in ASCE 7. The calculated wind pressure for Exposure Category C
based on the updated GRF results in only a 3% difference (reduction) for a
typical 100 ft structure compared to the previous method. The gust response
factor equations are discussed further in this section, and an example of
their application can be found in Appendix F. Further discussion on the
updated gust response factors is provided by Mara (2015).

2.1.5.1 Equations and Nomenclature  The gust response factor equations


presented in this section have been simplified from the complete gust
response factor equations described in Appendix F. That is, only the
background component of the dynamic response is considered, which is
likely sufficient for typical structures and span lengths. For tall or unique
structures, or for longer span lengths, designers are encouraged to consider
the complete gust response factors (including the resonance component)
provided in Appendix F.
The simplified structure and wire gust response factors, Gt and Gw,
respectively, are determined from the following equations:

 1 + 4.6 I z Bt 
Gt =   (2-4)
 1 + 6.1I z 
32 Electrical Transmission Line Structural Loading

 1 + 4.6 I z Bw 
Gw =   (2-5)
 1 + 6.1I z 

1
 33 6
in which I z = cexp   (2-6)
 z 
h

1
Bt =
0.56 zh (2-7)
1+
Ls

1 (2-8)
Bw =
0.8S
1+
Ls

where
Bt = Dimensionless response term corresponding to the
quasi-static background wind loading on the structure,
Bw = Dimensionless response term corresponding to the
quasi-static background wind loading on the wires,
cexp = Turbulence intensity constant, based on exposure (Table 2-3),
Iz = Turbulence intensity at effective height of the tower/struc-
ture or wire [Equation (2-6)],
Ls = Integral length scale of turbulence (ft) (Table 2-3),
S = Design wind span (ft) of the wires (conductors and ground
wires), and
zh = Two-thirds of the total height of the structure (ft) for the
calculation of Gt [Equation (2-4)], or effective height of the
wire (ft) for the calculation of Gw [Equation (2-5)].

Table 2-3.  Turbulence Parameters for Calculation of


Gust Response Factor by Exposure

Exposure cexp Ls (ft)

B 0.3 170

C 0.2 220

D 0.15 250
Weather-Related Loads 33

2.1.5.2 Gust Response Factor for Structure


The structure gust response factor, Gt, is used in Equation (2-1) for com-
puting the wind loads acting on transmission structures, structural com-
ponents, and on the insulator and hardware assemblies attached to the
structures. Gt , given by Equation (2-4), is a function of the exposure cate-
gory (defined in Section 2.1.4.1) and the effective height, zh.
The equation for Gt was developed from the complete equations dis-
cussed in Appendix F, and assumes the resonant component of the response
of the structure is negligible. The relationship for Gt as a function of total
structure height is plotted in Figure 2-7 for each exposure. In this plot, the
effective height of the tower is taken as 2/3 of the total height of the struc-
ture. As the resonant component is not included, Equation (2-4) yields
identical values for all structure types (e.g., self-supported latticed towers,
guyed towers, monopole structures, H-frame structures). This simplified
equation is applicable for most transmission structure types. For taller
structures, a more detailed analysis which considers the resonant compo-
nent is recommended.

Figure 2-7. Structure gust response factor.

2.1.5.3 Gust Response Factor for Wires


The wire gust response factor, Gw, is used in Equation (2-1) for comput-
ing the peak dynamic wind loads resulting from wind on the conductors
and overhead ground wires. It is given by Equation (2-5). Gw is a function
of exposure category (defined in Section 2.1.4.1), design wind span between
structures, and the effective height of the wires, zk.
Equation (2-5) and the curves for Gw in Figure 2-8 were developed from
the simplified equations discussed in this section and assume that the reso-
nant response of the wires to gusting wind is negligible. The resonant com-
ponent of the dynamic response is often quite small due to the lack of gust
correlation in the line direction and the significance of aerodynamic damp-
ing in the wires.
34 Electrical Transmission Line Structural Loading

(a)

(b)

(c)

Figure 2-8. (a) Wire gust response factor for Exposure B for different effective
heights, zh; (b) wire gust response factor for Exposure C for different effective
heights, zh; (c) wire gust response factor for Exposure D for different effective
heights, zh.

2.1.6 Force Coefficient


The force coefficient, Cf, in the wind force equation [Equation (2-1)]
accounts for the effects of member characteristics (e.g., shape, size, solidity,
Weather-Related Loads 35

shielding, orientation with respect to the wind, surface roughness) on the


resultant force. The force coefficient is the ratio of the resulting force per
unit area in the direction of the wind to the applied wind pressure. It can
also be referred to as the drag coefficient, pressure coefficient, or shape
factor.

2.1.6.1 Factors Influencing Force Coefficients  This section discusses


some of the important factors in the determination of the force coefficient
for a member or assembly of members. Additional theoretical background
can be found in Hoerner (1958), Davenport (1960), Sachs (1978), and Mehta
and Lou (1983).

2.1.6.1.1 Shape and Size  Member shapes fall into two general classifica-
tions: bluff and streamlined. The forces due to wind on a bluff structure
can be primarily attributed to the pressure distribution around the shape.
For streamlined shapes, such as airplane wings, friction accounts for the
majority of the drag force. Most buildings and engineering structures are
treated as bluff bodies (MacDonald 1975, Holmes 2001).
Bluff bodies can be considered to be divided into two classes: sharp-
edged and rounded. For sharp-edged members, such as rolled structural
shapes, the pressure distribution around the body remains relatively con-
stant for a given shape regardless of size or wind speed. These members
are often referred to as flat-sided members. A single force coefficient is
provided for flat-sided members. For rounded members, which are con-
sidered to be semi-aerodynamic, the pressure distribution around the body
varies with wind speed. Above a particular wind speed, referred to as the
critical wind speed, the negative pressure on the leeward side of the shape
decreases in magnitude. This causes a reduction in the overall force coef-
ficient of the member. The wind speed at which this change occurs is
dependent on the Reynolds number, which is a dimensionless ratio that
relates the inertia force (pressure) of the wind to its viscous force (friction).
The equation for the Reynolds number is given as

Re = 9350 K z VMRI(ds/12) (2-9)

where
Re = Reynolds number, referenced at standard atmosphere,
Kz = Terrain factor at height z above ground (Table 2-2),
VMRI = Basic design wind speed (mph) (Section 2.1), and
ds = Diameter of the conductor or ground wire, or the width of the
structural shape normal to the wind direction (in inches).
36 Electrical Transmission Line Structural Loading

Ice accretion on wires and structural members changes the force coef-
ficient for these components; refer to Section 2.3.4.3, Section 2.3.5.2, and
Appendix G.

2.1.6.1.2 Aspect Ratio  The ratio of the length of a member to its diameter (or
width) is known as the aspect ratio. Short members have lower force coef-
ficients than long members of the same shape. The force coefficients given
in Section 2.1.6.2 are applicable to members with aspect ratios greater than
40, which is typical of most transmission line structures. Correction factors
for members with aspect ratios less than 40 are given in Appendix G.

2.1.6.1.3 Solidity Ratio  An important parameter that influences the force


coefficient for latticed truss structures is the solidity ratio of the faces. The
force coefficient for the total structure is dependent on the airflow resis-
tance of individual members and on the airflow patterns around the mem-
bers. The force coefficients shown in Section 2.1.6.2 are a function of solidity
ratio, Φ, which is defined as
Am
Φ= (2-10)
Ao

where Am is the area of all members in the windward face of the struc-
ture (net area), and Ao is the area of the outline of the windward face of the
structure (gross area).
The solidity ratio of each discrete panel in the transverse and longitudi-
nal faces should be used for the determination of wind loads.

2.1.6.1.4 Shielding  If an upstream body (e.g., structural element, compo-


nent, wire) protects a downstream body from the wind, either partially or
completely, it is referred to as shielding. The amount of shielding experi-
enced by the downwind body is influenced by the solidity ratio of the
upstream body, the spacing between the bodies (referred to as separation
distance), and the yaw angle of the wind. The shielding factor is defined
as the ratio of the force coefficient of the shielded body to the force coeffi-
cient of the unshielded body.
Shielding factors for trusses and frames are available in some codes and
literature (i.e., NRCC 2010, Davenport 1960, SIA 1956). The expressions for
the force coefficients of square and triangular latticed truss structures in
Table 2-4 already include the effect of shielding on the downwind faces of
the truss structures.
There are typically no shielding effects between individual poles of
H-frames or multi-pole structures due to the low ratio of member diameter
to separation distance and the small range of yaw angles for which a down-
wind pole would be shielded. For similar reasons, shielding effects are not
Weather-Related Loads 37

typically considered for wires (e.g., bundled conductors, adjacent phases/


poles), or transmission structures which are in line with one another.

2.1.6.1.5 Yawed Wind  The term yawed wind is used to describe wind for
which the angle of incidence with the face or shape is not perpendicular.
The yaw angle, Ψ, is measured in the horizontal plane and is referenced as
0° for wind perpendicular to the conductors. Figure 2-9 shows an example
of yawed wind and the resultant force directions. Transverse loads act per-
pendicular to the direction of the transmission line, while longitudinal
loads act parallel to the direction of the transmission line. Expressions for
yawed wind on wires and structures are provided in Section 2.1.6.2.

Figure 2-9. Yawed wind on a transmission line.

2.1.6.2 Recommended Force Coefficients  The following sections describe


the force coefficients recommended by this manual for various components
of a transmission system. Other force coefficients can be used where
justified by experimental data (i.e., wind tunnel testing). Additional
background information on force coefficients can be found in Appendix G.

2.1.6.2.1 Conductors and Ground Wires  Many designers currently use a


force coefficient of 1.0 for conductors and ground wires, as indicated in
NESC Rule 251 (NESC 2017). Values from wind tunnel testing, which are
described in detail in Appendix G, range from 0.7 to 1.35. These data exhibit
large variations in the wire force coefficient over a wide range of Reynolds
numbers. Unless more definitive data based on wind force measurements
38 Electrical Transmission Line Structural Loading

are available (e.g., wind tunnel testing), a constant force coefficient value
of

Cf = 1.0 (2-11)

is recommended for single or bundled conductors and for ground wires.


Smaller wire sizes typically have a higher force coefficient (Appendix G).
Note that if a reduced value of Cf is used on bare wires based on wind tun-
nel data, instances of wind loading for ice-covered wires should revert to
a value of 1.0.
Equation (2-1) may be multiplied by cos2Ψ to account for yawed wind
on conductors and ground wires, in which Ψ is the yaw angle (Figure 2-9).
The equation for wind force on conductors for yawed winds is given in
Equation (2-12). For all yaw angles, the effective force calculated by Equa-
tion (2-12) is perpendicular to the conductor or ground wire.

F = QKzKztVMRI2GwCfAcos2Ψ (2-12)

2.1.6.2.2 Latticed Truss Structures  This manual recommends that force coef-
ficients for square-section and triangular-section (in plan view) latticed
truss structures be determined from ASCE 7-16 unless other requirements
dictate the design. The relevant force coefficients are shown in Table 2-4
and Figure 2-10, as obtained from ASCE 7-16. These force coefficients
account for both the windward and leeward faces, including the shielding
of the leeward face by members in the windward face. Therefore, the force
coefficients are multiplied by the projected area of one tower face only.
The ASCE 7-16 force coefficients for square-section and triangular-section
latticed truss structures having flat-sided members are given in Table 2-4.
The force coefficients given in this table for square-section structures may
also be used for rectangular-section structures. For towers with round-
section member shapes, the force coefficients are determined by multiply-
ing the value calculated from Table 2-4 by Equation (2-13). Note that the
correction factor for rounded members as calculated by Equation (2-13) has
a limit of 1.0. The relationship of the force coefficient and solidity ratio is
plotted for each cross-section type in Figure 2-10. Caution should be used
when applying the force coefficients provided in Table 2-4 to sections of
unique geometry, as the equations are primarily based on aerodynamic
results for typical sections; additional discussion is given by Mara (2014).
Weather-Related Loads 39

(a)

(b)

Figure 2-10. Force coefficients for (a) square; and (b) triangular cross sections.
40 Electrical Transmission Line Structural Loading

Table 2-4.  Force Coefficients, Cf, for Normal Wind on


Latticed Truss Structures Having Flat-Sided Members

Tower cross-section Cf

Square 4.0Φ2 − 5.9Φ + 4.0

Triangular 3.4Φ2 − 4.7Φ + 3.4


Source: Adapted from ASCE 7-16 (ASCE 2017).

Correction factor (for round members)


= 0.51Φ2 + 0.57, but not greater than 1.0 (2-13)

2.1.6.2.3 Latticed Truss Structures: Transverse, Longitudinal, and Yawed Wind 


Various methods are available for the calculation of wind loads on latticed
truss structures. This manual presents two methods for the calculation of
transverse, longitudinal, and yawed wind loads. The fundamentals of each
method are presented in this section, along with the assumptions and key
points that should be considered by the designer should either method be
used. It is recognized that the two methods presented do not provide equal
results; however, either method may be used by the designer provided the
parameters are appropriately addressed. Either method may provide a
greater wind load than the other, which is dependent on the selected
parameters and application. The method resulting in the greater load does
not imply correctness or that it should be considered the default method.
It is the responsibility of the designer to provide justification for the method
and parameters used. Other methods may be used provided good engi-
neering judgment is applied.
The first method is based on identifying the aerodynamic properties (net
area and force coefficient) of the transverse and longitudinal faces of the
tower, typically in multiple sections over the height. This is commonly
referred to as the “Wind on Face” method.
The second method is based on calculating the wind load on each indi-
vidual member of the tower. This is commonly referred to as the “Wind on
Member” method.

Method 1: Wind on Face  This method starts with the segmenting of the
tower geometry at a reasonable number of heights and calculating the net
area and solidity ratio for each segment (or panel). The force coefficient for
each panel is then calculated based on Table 2-4. The resultant wind force
is then calculated for each panel following Equation (2-14a). An expression
Weather-Related Loads 41

similar to Equation (2-14a) is currently used in the International Electro-


technical Commission Standard 60826 (2017).

Fd = QKzKztVMRI2GtδΨ(CftAmtcos2Ψ + CflAmlsin2Ψ) (2-14a)

where
Fd = Force in the direction of the wind [lb (N)],
Amt = Area of all members in the face of the structure that is parallel
to the line [ft2 (m2)],
Aml = Area of all members in the face of the structure that is perpen-
dicular to the line [ft2 (m2)],
Cft = Force coefficient associated with the face of the structure that
is parallel to the line,
Cfl = Force coefficient associated with the face of the structure that
is perpendicular to the line,
δΨ = Wind angle magnification factor equal to [1 + 0.2sin2(2Ψ)]
The other variables are as defined previously.

The resultant wind force is then decomposed into forces in the trans-
verse and longitudinal directions, Ft and Fl, respectively, to assist with the
design of each direction.

Ft = FdcosΨ (2-14b)

Fl = FdsinΨ (2-14c)

The presentation of this method differs from that recommended in the


previous edition of this manual. While little difference exists between this
method and that previously used for sections which are symmetric in plan
view (i.e., square), the aerodynamics of portions of the structure which are
not symmetric in plan view (i.e., cross-arm, bridge) are better captured
through the current method.
As presented, the “Wind on Face” method contains some assumptions.
Most notably these assumptions include
• The selection of tower face segments should represent the different
structural member panel profiles for calculation of Amt, Aml, and the
resulting solidity ratios.
• The force coefficients used (Table 2-4) are based on those available
in the literature for symmetric (in plan) square and triangular
sections. The design under consideration may not reflect these
characteristics. However, if the extents of the face (and thus the
solidity ratio) are properly selected, this method should provide
acceptable values of the force coefficients.
42 Electrical Transmission Line Structural Loading

• The equation is based on the assumption that the maximum wind


load on the structure occurs for a yaw angle of 45°. This results
from the wind angle magnification factor δΨ in Equation (2-14a).
Wind tunnel tests on models of square sections (BEAIRA 1935,
Bayar 1986) and a cross-arm section (Mara et al. 2010) have demon-
strated that the maximum effective wind loads are created by
winds at yaw angles ranging from 30° to 45°. The magnitude of
these increases will vary with tower geometry.

Method 2: Wind on Member  The overall wind loads on a tower are calcu-
lated based on developing wind forces acting on each member and distrib-
uting the load among the joints at the end of each member. The wind loads
on each member are then summed along the height of the tower. This
method is most conveniently applied using computational techniques, as
there are significantly more calculations to be performed than with the
“Wind on Face” method.
This method begins with the application of wind pressure on each mem-
ber based on the geometric relationship between the global wind velocity
vector, the member joint-to-joint orientation with respect to the global axis,
and the cross-section orientation of the member with respect to its local
axis. The calculated force, Fm, is in the plane formed by the wind velocity
vector (which acts in the horizontal plane) and the global member orienta-
tion. The magnitude of the wind load on each member is based on Equa-
tion (2-15)

Fm = QKz,mKzt,mVMRI2GtCf,mAmocos2α (2-15)

where
Fm = Force acting perpendicular to a member, in the plane formed
by the horizontal wind velocity vector and the global mem-
ber orientation (lb);
Kz,m = Velocity pressure exposure coefficient for the member, based
on the average height of the member;
Kzt,m = Topographic factor for the member, based on the average
height of the member;
Cf,m = Force coefficient of the shape of member;
Amo = Maximum exposed wind area on the member, based on the
length of the member and the maximum dimension exposed
to wind (ft2);
α = Incidence angle, defined as the angle formed by horizontal
wind velocity vector and the plane perpendicular to the
global member orientation (degrees).
The other variables are as defined previously.
Weather-Related Loads 43

The calculation of the incidence angle, α, for each member at each yawed
wind direction, Ψ, requires the global member orientation to be determined
from the three-dimensional geometry of the structure. The selection of the
force coefficient of the member, Cf,m, should be representative of the type
of member under consideration. Force coefficient values for different mem-
ber shapes are provided in Appendix G.
The force acting perpendicular to the member, Fm, is then resolved into
the horizontal plane based on the global member orientation. The resulting
horizontal force is then subsequently resolved into the transverse and lon-
gitudinal directions for the calculation of overall wind loads. The loads
resulting from wind on each member are then summed and determined at
each joint along the height of the tower.
As presented, the “Wind on Member” method contains some assump-
tions. Most notably these assumptions include
• The selected force coefficient, Cf,m, should reflect the type of mem-
ber cross-section and level of inclusion of structural and non-struc-
tural members contained in the model.
• This method assumes that the force coefficient selected for the
member is consistent among all wind directions and member
orientations. A single, selected value of Cf,m will neglect the true
member force coefficient as it appears to the wind for all incidence
angles other than that of the profile orientation based on the
respective Cf,m selected, and therefore a conservative value should
be used.
• Applying this method to members whose yaw angles are parallel to
the global member orientation (i.e., horizontal members in line with
the wind) will produce zero load.
• This method assumes that no shielding or flow acceleration around
members occurs. These are both complex mechanisms, and the
designer should be comfortable with the underlying aerodynamics
if shielding effects are to be considered for a particular member or
members. Additional discussion of shielding can be found in
Section 2.1.6.1.4.
If the assumptions are applied, this method should provide conservative
estimates of the yawed angle wind loads.

2.1.6.2.4 Pole Structures  The total perpendicular or yawed wind force on


single-shaft and H-frame structures is the sum of the wind forces on the
individual members within the structure. Typically, transmission pole
shafts and closed cross-sectional shapes exceed 1 ft in diameter, which
results in a Reynolds number greater than 6 × 105 based on the design wind
speeds specified in Figure 2-1.
44 Electrical Transmission Line Structural Loading

Table 2-5.  Force Coefficients, Cf, for Members of Pole Structures.

Member shape Cf Adapted from


Circular 0.9 ASCE 7-16
16-sided polygonal 0.9 James (1976)
12-sided polygonal 1.0 James (1976)
8-sided polygonal 1.4 ASCE 7-16, James (1976)
6-sided polygonal 1.4 ASCE 7-16
Square, rectangle 2.0 ASCE 7-16

Surface roughness (e.g., rough for wood, smooth for steel) will influence
the force coefficients for these shapes. Attachments on pole structures, such
as steps, ladders, arms, and brackets, will also influence the force coeffi-
cients. The effects of attachments and surface conditions can be significant
on streamlined shapes such as circular members.
Table 2-5 lists recommended force coefficients for structural shapes com-
monly used for transmission pole structures. These coefficients are based
on research by James (1976) and on values given in ASCE 7-16. The recom-
mended force coefficients include the effect of typical surface roughness
and attachments, such as steps, ladders, and brackets. For example, the
force coefficient for a circular member is based on ASCE 7-16 assuming a
rough surface. This accounts for the surface condition of a wood pole or
typical steel pole attachments. For 12-sided and 16-sided polygonal shapes,
the corner radius ratio term from James (1976) has been omitted to account
for the effect of typical attachments.
In certain cases, it may be appropriate to select force coefficients greater
than those listed in Table 2-5. Appendix G provides additional force coef-
ficients for various shapes. The use of these or other values should be based
on either design experience or research results.

2.1.6.2.5 Other Members and Components  Appendix G also lists force coef-
ficients for structural shapes based on Reynolds number, corner radius,
and yaw angle. The effects of steps, ladders, arms, brackets, and other
attachments are not included in the values shown in Appendix G.

2.1.7 Topographic Effects


The wind speeds which transmission line structures experience can be
significantly influenced by topography. Specific recommendations on the
treatment of some topographic effects are beyond the scope of this docu-
Weather-Related Loads 45

ment. The designer may benefit from the advice of a meteorologist or wind
engineer in situations where topographic effects are expected to be severe.
Guidelines on the effects of isolated hills and ridges on wind speeds are
available (e.g., ASCE 7-16). In addition, extensive field programs and
research have been devoted to the subject of boundary layer flow over hills
and complex terrain (Walmsley et al. 1986, Taylor et al. 1987, Miller and
Davenport 1998). Examples of topographic effects include channeling of
wind, flow around mountains and hills, and flow through canyons and
valleys. Note that some of these instances are treated as special wind
regions in Figure 2-1.

2.1.7.1 Channeling of Wind  This effect occurs where there is a natural


flow of air from an unrestricted area through a restricted area, such as a
mountain pass. As air is channeled into an opening or canyon, it accelerates
due to the Venturi effect. This type of wind is often referred to as a local
canyon wind. The wind velocity through a canyon may be as much as
double that in the unrestricted areas on each side. If this condition arises
along the right-of-way, the design loads should be adjusted accordingly.
Buildings may create a similar channeling effect to mountains. Generally
speaking, buildings would not be a major influence on a transmission line.
They could, however, alter the wind loading on one or two structures and
the designer should be aware that wind channeling generated by adjacent
buildings could occur in certain locations along the line.

2.1.7.2 Mountains  Wind tunnel tests (Britter et al. 1981, Arya et al. 1987,
Snyder and Britter 1987, Gong and Ibbetson 1989, Finnigan et al. 1990) and
field experiments (Coppin et al. 1994) suggest that wind speed can increase
in localized areas of mountains on the windward side as well as on the
leeward side (Armitt et al. 1975). When the wind is blowing normal to a
mountain ridge, the air compresses as it moves up the windward side. With
any opening in the ridge, the compressed air is released and accelerates as
in the case of local canyon winds.
With the appropriate combination of pressure and temperature, the
wind passing over a mountain ridge accelerates on the leeward side. Accel-
erated winds of this type are sometimes called Santa Ana, Chinook, stand-
ing wave, or downslope winds. Several areas in the United States
experience downslope winds due to their proximity to mountain ridges.

2.1.7.3 Wind Speed-up over Hills, Ridges, and Escarpments  Wind


speed-up due to flow over hills and escarpments is addressed in ASCE
7-16, as illustrated in Figure 2-11. The provisions apply to isolated hills or
escarpments located in Exposure Categories B, C, or D. The topographic
feature [two-dimensional ridge or escarpment, or three-dimensional axis-
symmetric hill] is described by two parameters, H and Lh, as indicated in
46 Electrical Transmission Line Structural Loading

Figure 2-11. Topographic factor after ASCE 7-16.


Source: ASCE (2017).
Weather-Related Loads 47

Figure 2-11 (Continued). Topographic factor after ASCE 7-16.


Source: ASCE (2017).
48 Electrical Transmission Line Structural Loading

Figure 2-11; H is the height of the hill, or difference in elevation between


the crest and that of the upwind terrain, and Lh is the distance upwind of
the crest to where the ground elevation is equal to half the height of the hill.
The topographic features may be considered in the design and location
where the upwind terrain is free of such topographic features for a distance
equal to 100H or 2 miles (3.2 km), whichever is less. The effect of wind
speed-up does not need to be considered when H/Lh < 0.2, or when
• H < 60 ft (18 m) for Exposure B
• H < 15 ft (4.5 m) for Exposures C and D.
To account for the wind speed-up over isolated hills and escarpments
that constitute abrupt changes in the general topography, a topographic
factor, Kzt, may be applied to the transmission structures sited on the upper
half of hills and ridges or near the edges of escarpments. The expression
for Kzt is given in Equation (2-16).

Kzt = (1 + K1K2K3)2 (2-16)

Definitions of the multipliers K1, K2, and K3 are given in Figure 2-11. The
multipliers are based on the assumption that the wind approaches the hill
along the direction of maximum slope, causing the greatest speed-up near
the crest. The value of Kzt should not be less than 1.0. It is not the intent of
this section to address the case of wind flow over complex terrain (such as
mountainous terrain, or non-isolated hills or escarpments) for which engi-
neering judgment, expert advice, or wind tunnel testing may be required.

2.1.7.4 Canyons and Valleys  Transmission lines may be subject to high


winds passing through canyons, from cool air masses spilling over a ridge,
or from general winds moving through a valley. Air masses spilling over
into a valley can be several miles in width and may reach velocities in
excess of 100 mph (160 km/h). This kind of event can occur several miles
from a mountain range.

2.1.8 Application of Wind Loads to Latticed Towers


There is no standard procedure for the application of the wind forces
determined from Equation (2-1) or (2-14) to the panel points of a latticed
tower. Typically, the designer will follow the procedures specified by the
individual utility. For example, some utilities may distribute the wind
forces to the windward panel points, whereas others may distribute the
wind forces to all panel points at an elevation (i.e., windward and leeward
panel points). However, all utilities generally distribute wind forces to the
respective member connecting joints as concentrated vector loads. A few
Weather-Related Loads 49

key points should be considered when applying the calculated wind forces
on different types of structures.
The wind forces determined by Equation (2-1) or (2-14), using the rec-
ommended force coefficients in this manual, account for loads accumulated
by both the windward and leeward tower faces (including shielding
effects). Therefore, the wind forces calculated on a single-body latticed
truss system, such as a vertical double-circuit self-supported structure, can
be distributed to the panel points of the structure without further
consideration.
For separated latticed truss systems and individual tubular shaft mem-
bers, such as a guyed portal or an H-frame structure, the windward faces
of each mast should be considered as being individually exposed to the
calculated wind force determined from Equation (2-1) using the appropri-
ate force coefficients. The wind forces can then be distributed along the
structure panel points according to the criteria specified by the utility.
Other locations on a structure may need to be reviewed where physically
separated latticed truss systems or tubular shaft members are used.
Longitudinal winds may also result in significant structural loading.
This case should be considered in the design of the structure.

2.2 HIGH-INTENSITY WINDS

Tornadoes and downbursts are the high-intensity winds (HIWs) dis-


cussed in this section. HIWs are generally the result of intense, localized
thermal activity that frequently accompanies a thunderstorm or squall line.
These HIWs are commonly narrow-front winds with speeds greater than
the sustained, broad-front, synoptic winds described in Section 2.1. HIWs
do not follow the pattern and characteristics of extreme winds from which
the mathematics of gust response factors in Section 2.1 were developed.
Analyses of line failures in several countries have identified HIW events
as the leading cause of transmission line failures. It is possible to apply
rational measures to transmission line design to increase the reliability of
transmission lines impacted by the majority of HIWs in the absence of
windborne debris.

2.2.1 Downbursts
2.2.1.1 Background  A downburst is defined as an intense downdraft of
air that induces high-velocity winds in all directions when striking the
ground. Fujita (1985) defined a downburst as a mass of cold and moist air
that drops suddenly from the thunderstorm cloud base, impinging on the
ground surface and then transferring horizontally. The practical diameter
50 Electrical Transmission Line Structural Loading

Figure 2-12. Downburst polar coordinates with respect to the tower of interest.

of this initial cold air jet can vary between 2,000 ft (600 m) and 5,600 ft
(1,700 m) as indicated by Hjelmfelt (1988).
The distribution and magnitude of the forces acting on the tower and
conductors depend significantly on the downburst characteristics, which
are the jet velocity (VJ), the jet diameter (DJ), and the location of the down-
burst center relative to the tower (represented by the polar coordinates R
and Ψ shown in Figure 2-12).

2.2.1.2 Downburst Wind Field  The downburst outflow velocity at an


arbitrary point in space has two components: a radial component VRD and
a vertical component VVR. Using a computational fluid dynamics simulation
of the spatial and time variations of the wind field associated with
downbursts (Hangan et al. 2003, Shehata and El Damatty 2007) concluded
that the vertical velocity component has a negligible effect on the structural
response in comparison to the radial velocity component. As such, the
current section focuses on describing the wind field associated with the
radial velocity. Figure 2-13 illustrates the variation with height of the radial
velocity, normalized with respect to the jet velocity (El Damatty and
Elawady 2015). The maximum velocity occurs at an R/DJ value of
approximately 1.3. The velocity profile in Figure 2-13 is provided in terms
Weather-Related Loads 51

Figure 2-13. Vertical profile of the radial outflow wind associated with a
downburst considering DJ = 1,640 ft (500 m).

of downburst jet velocity, VJ, which represents the speed at which the
downburst event impinges on the surface of the earth. In view of the
information found in the literature (e.g., Fujita 1985, Orwig and Schroeder
2007, CIGRÉ 2009), a value for VJ of 112 to 157 mph (50 to 70 m/s) is
recommended.
The span reduction factor commonly used to adjust the loads applied
on the conductor spans is closer to unity when compared with that in syn-
optic winds (Holmes et al. 2008, Aboshosha and El Damatty 2013). One
approach for simulating critical downburst load cases for transmission line
structures is provided in Appendix K.

2.2.2 Tornadoes
2.2.2.1 Background  The usual perception of a tornado is of an over­
whelming event destroying all in its path and defying resistance. Although
most tornadoes are capable of causing severe damage to houses, mobile
homes, and automobiles, most engineered structures in the vicinity of
tornadoes survive without major damage. Transmission line structures can
be designed with sufficient strength to resist wind loads incurred in a
majority of tornado events (F0–F2). However, designing for severe
tornadoes (i.e., F3 to F5) may be prohibitive due to significantly higher cost.
For these severe tornadoes, the focus of the designer changes from resisting
the HIWs to failure containment.
52 Electrical Transmission Line Structural Loading

Tornadoes occur on most subtropical and temperate landmasses around


the world. On average, 800 to 1,000 tornadoes occur each year in the con-
tiguous United States, and the activity zone extends well into Canada. The
total number of reported tornadoes in 1° squares of latitude and longitude
for a 30-year period (1950 to 1980) is shown in Figure 2-14. A 1° square
contains about 4,000 mi2 (10,000 km2). More recent information on tornado
characteristics, occurrence, and forecasting can be obtained at the online
Storm Prediction Center, which is maintained by the National Oceanic and
Atmospheric Administration (NOAA)/National Weather Service.
Fujita and Pearson (1973) have developed a rating (F0 to F5) to catego-
rize tornadoes by their intensity and size. This method assigns a numerical
value of the F-scale to each tornado based on the appearance and extent of
damage. The F-Scale and the associated path length, path width, and wind
speed are shown in Table 2-6. The wind speed provided the fastest
quarter-mile speed, assumed at 16 to 33 ft (5 to 10 m) above the ground
level, and also as 3-second gust wind speed. Also provided are the number
of tornadoes, percentage, and cumulative percentage for each scale. It is
noted that 86% of the tornadoes are assigned to the scale of F2 or smaller;
the F2 rating corresponding to a 3-second gust wind speed of approxi-
mately 160 mph (70 m/s) or less.
The probability of a tornado strike at a given point is very small (McDon-
ald 1983), even in areas of tornado prevalence. However, the probability of
a transmission line being crossed by a tornado is significant (Twisdale

Figure 2-14. Total number of reported tornadoes during a 30-year period.


Source: Tecson et al. (1979).
Table 2-6.  Tornado Wind Speed, Path Length, Path Width, and Frequencies for F-Scale in the United States

Fujita Scale
Fastest quarter-mile 3-second gust Tornado frequencies (1916–1978)*
Scale
Path length Path width Number of Cumulative
Wind speed Wind speed Percentage
tornadoes percentage
72 (mph) <1.0 (mi) <50 (ft)
F0 45–78 (mph) 5,718 22.9 22.9
32.2 (m/sec) 1.61 (km) 15.2 (m)
73–112 (mph) 1.1–3.1 (mi) 51–170 (ft)
F1 79–117 (mph) 8,645 34.7 57.6
32.6–50 (m/s) 1.8–5.0 (km) 15.5–52 (m)
113–157 (mph) 3.2–9.9 (mi) 171–530 (ft)
F2 118–161 (mph) 7,102 28.5 86.1
50.5–70.2 (m/s) 5.1–15.9 (km) 52.1–161.5 (m)
158–206 (mph) 10–31 (mi) 531–1,670 (ft)
F3 162–209 (mph) 2,665 10.7 96.8
70.6–92.1 (m/s) 16–50 (km) 1,61.8–509 (m)
207–260 (mph) 32–99 (mi) 1,671–4,750 (ft)
F4 210–261 (mph) 673 2.7 99.5
92.5–116.2 (m/s) 51–159 (km) 509.3–1,447 (m)
261–318 (mph) 100–315 (mi) 4,751–6,000 (ft)
F5 262–317 (mph) 127 0.5 100
116.7–142.2 (m/s) 160–507 (km) 1,448–1,829 (m)
*Source: Tecson et al. (1979)
54 Electrical Transmission Line Structural Loading

1982). The fact that the width of the path is very narrow for most tornadoes,
however, makes it possible to improve transmission line resistance to the
majority of tornadoes. Almost all tornadoes can engulf a house or small
structure, but very few have a width that will load the full span of a trans-
mission line, with the exception of spans less than perhaps 500 to 600 ft
(150 to 200 m).

2.2.2.2 Tornado Wind Field  The wind pattern within a tornado is


composed of circular winds combined with translation, the highest
velocities being where the rotary and translation components add together
(Abbey 1976; Mehta et al. 1976; Minor et al. 1977; Wen and Chu 1973;
Hangan and Kim 2008; Hamada et al. 2010; Hamada and El Damatty 2011,
2015). An idealized pattern of tornado wind velocities is shown in
Figure 2-15. The tornado wind velocity at a certain point in space has three
components: tangential velocity, radial velocity, and uplift (axial) velocity.
Field data for the 1998 Spencer, South Dakota, F4 tornado and for the
1999 Mulhall, Oklahoma, F4 tornado were used to validate the numerical
simulations of F4 and F2 tornadoes (Hangan and Kim 2008, Hamada et al.
2010) that are discussed subsequently and in Appendix K.
Idealized vertical profiles for the axisymmetric average of tangential,
radial, and vertical velocity components below 330 ft (100 m) are provided
in Figures 2-16 to 2-18 at three radial distances, r, where r is the distance

Figure 2-15. Hypothetical pattern of tornado wind velocities and directions.


Weather-Related Loads 55

measured from the center of a tornado. The radial velocity changes direc-
tion with height, where negative values act in an inward direction and
positive values act in an outward direction. A positive value for the axial
or vertical component indicates an upward velocity. One approach for
simulating critical tornado load cases for transmission line structures is
provided in Appendix K.

Figure 2-16. Idealized vertical profiles of tangential velocity component for three
radial distances from F2 tornado center.

Figure 2-17. Idealized vertical profiles of radial velocity component for three
radial distances from F2 tornado center.
56 Electrical Transmission Line Structural Loading

Figure 2-18. Idealized vertical profiles of vertical velocity component for three
radial distances from F2 tornado center.

2.3 ICE AND WIND LOADING

2.3.1 Introduction
Ice accretion on a transmission line is often a governing loading criterion
in structure design. In addition to imposing substantial vertical loads on
the structural system, ice buildup on the wires and conductors presents a
greater projected area exposed to the wind and affects the force coefficient.
The resultant load on the wires and conductors may cause significantly
greater wire tensions compared to the bare conductor conditions. Meteo-
rological data suggest, and a survey of utility practice (ASCE 1982) con-
firms, that it is prudent to include concurrent ice and wind loadings in the
load criteria of transmission structure designs throughout most of the
United States.
The following discussion provides general guidance for the selection of
ice with concurrent wind loads. Where more detailed icing data have been
compiled for a service area, those data should take precedence over the
information in this manual. Electric utilities are urged to develop concur-
rent ice and wind loading criteria specifically for their service regions
based on historical data.

2.3.2 Categories of Icing


Ice can be classified by either its method of formation or its physical
characteristics. Precipitation icing from freezing rain or freezing drizzle is
the most common icing mechanism.
Weather-Related Loads 57

The inglaze ice that forms in these conditions is usually clear but may
also be translucent because of included air bubbles.
In-cloud icing is caused by supercooled cloud droplets, carried by the
wind, colliding with a surface. The ice that forms ranges from hard, clear
glaze to softer, lower-density white rime ice containing entrapped air.
In-cloud icing may occur in regions with level terrain, but is more fre-
quently associated with mountainous areas, occurring on both exposed
summits and upslopes.
Snow, both wet and dry, may adhere to wires by capillary forces, freez-
ing and sintering, forming a cylindrical sleeve around the wire. The density
of accreted snow depends on the wind speed and wetness of the snow.
Hoarfrost is an accumulation of ice crystals formed by the direct deposi-
tion of water vapor from the air onto a structure. The amount of ice accreted
by vapor deposition does not impose significant loads on structures.
It is important that the transmission line engineer be aware of the icing
conditions (i.e., freezing precipitation, in-cloud icing, or sticky snow) that
may occur along the route of a proposed transmission line. Ice accretions
produced by freezing rain rarely exceed a thickness of a few inches,
whereas lower-density accretions due to in-cloud icing and sticky snow
can build to thicknesses of a foot or greater. Furthermore, in-cloud icing
can produce significant unbalanced loadings between adjacent spans with
different wind exposures. The designer would benefit from the advice of a
meteorologist in regions where in-cloud icing may be severe.
Section H.1 in Appendix H provides additional information on the mete-
orological conditions that are associated with the various types of icing and
properties of the ice accretions.

2.3.3 Design Assumptions for Ice Loading


The four categories of icing (glaze, in-cloud, snow, and hoarfrost) cover
the spectrum of the icing phenomenon. The distinctions made by definition
of each category may not be identifiable in practice. There can be an over-
lap of more than one type of icing condition, such as snow and freezing
rain, or in-cloud icing and freezing drizzle. In specifying ice loadings, the
accretion density should be noted and is typically assumed to be uniform
with thickness.
For simplicity, the design ice thickness is specified as an equivalent uni-
form radial thickness over the length of the wire. However, natural ice
accretions may be uniform, elliptical, crescent-shaped, pennant-shaped, or
have attached icicles.

2.3.4 Ice Accretion on Wires Due to Freezing Rain


2.3.4.1 Historical Ice Data  As weather stations do not collect data on ice
thickness, ice accretion models based on meteorological data are often used
58 Electrical Transmission Line Structural Loading

to estimate ice thickness. Where modeled or measured ice thickness and


concurrent wind speed data are available, the calculated nominal ice
thickness is defined as t (which is t100 or tMRI). The concurrent wind speed
is used to determine the concurrent transverse wind load that is combined
with the vertical load due to the weight of the ice.

2.3.4.2 Using Ice Accretion Maps  In areas where local historical icing
data are not available, the glaze ice maps given in Figures 2-19 through 2-23
can be used with some limitations. These maps show 100-year MRI ice
thicknesses due to freezing precipitation with concurrent 3-second gust
wind speeds V3-sec at 33 ft (10 m) above ground for the continental United
States and Alaska. These maps are revised from the maps in the previous
edition of this manual (ASCE 2010a) using data from weather stations in
Canada near the US border and using a 100-year MRI instead of a 50-year
MRI. The stations used to map Alaska are shown in Figure 2-23, and
stations used to map the lower 48 states are shown in Appendix H, Figure
H-1. The glaze ice thicknesses shown in these figures do not include
in-cloud icing or sticky snow accretions, which are caused by meteorological
conditions that may produce significantly different loading patterns (see
Appendix H, Section H.5).
Additional maps showing the 50-year and 300-year MRI ice thicknesses
due to freezing precipitation with concurrent 3-second gust wind speeds
are provided in Appendix L. The mapped ice thicknesses and wind speeds
are based on Exposure C, but should also be used for Exposures B and D.
The amount of ice that accretes on a wire depends on the wind speed at
the wire height. Design thicknesses of glaze ice tz for heights z above
ground can be obtained from Equations (2-17a) and (2-17b).

 z 
0.10

tz = tMRI   for 0 ft < z < 900 ft (2-17a)


 33 

 z 
0.10

tz = tMRI   for 0 m < z < 275 m (2-17b)


 10 

where
tMRI = Nominal ice thickness (e.g., t100 for 100-year MRI),
tz = Design ice thickness at height z, and
z = Height above ground (ft in customary units [Equa-
tion (2-17a)]; m in metric units [Equation (2-17b)]).

The exponent in Equation (2-17) is based on a power law increase of


wind speed with height for Exposure Category C (open country terrain).
Weather-Related Loads 59

(a)
Figure 2-19. 100-year MRI radial ice thickness (in.) from freezing rain with
concurrent gust wind speeds (mph) at 33 ft (10 m) aboveground in Exposure C
for the (a) western; and (b) eastern United States.

At sites that tend to be windy or where the wind speed increases rapidly
with height, the ice thickness gradient will be more pronounced than indi-
cated by Equation (2-17). The concurrent gust wind pressure is also
increased with height above ground using Equation (2-3) and the power
law exponents indicated in Table 2-1. Ice thicknesses on a ridge, hill, or
60 Electrical Transmission Line Structural Loading

(b)
Figure 2-19. (Continued) 100-year MRI radial ice thickness (in.) from freezing
rain with concurrent gust wind speeds (mph) at 33 ft (10 m) above­ground in
Exposure C for the (a) western; and (b) eastern United States.

escarpment will be greater than those in level terrain due to wind speed-up
effects. The topographic factor for the ice thickness on isolated ridges, hills,
or escarpments is Kzt0.35, where Kzt is obtained from Equation (2-16). It
should be noted that as ice thickness and concurrent wind are affected by
height above ground and topography, there are additional uncertainties
involved with this process than suggested by the preceding calculations.
Weather-Related Loads 61

Figure 2-20. 100-year MRI radial ice thickness (in.) from freezing rain with
concurrent gust wind speeds (mph) at 33 ft (10 m) aboveground in Exposure C:
Puget Sound detail.

For areas not covered by Figures 2-19 though 2-23 and areas where
in-cloud icing or sticky snow is the most severe icing mechanism, other
sources of information must be consulted to determine design ice thick-
nesses; refer to Appendix H, Sections H.4 and H.5, for additional informa-
tion. Figures 2-19 through 2-23 represent glaze ice thickness values at single
points, and do not include spatial effects (refer to Appendix B).
In areas where little information on ice loads is available, it is recom-
mended that a meteorologist familiar with atmospheric icing be consulted.
Factors to be kept in mind include that taller structures may accrete more
ice due to higher winds and colder temperatures aloft, and that influences
of elevation, complex relief, proximity to water, and potential for unbal-
anced loading are significant.

2.3.4.3 Ice with Concurrent Wind Loading  The ice thicknesses due to
freezing rain in Sections 2.3.4.1 and 2.3.4.2 are uniform radial glaze ice
62 Electrical Transmission Line Structural Loading

Figure 2-21. 100-year MRI radial ice thickness (in.) from freezing rain with
concurrent gust wind speeds (mph) at 33 ft (10 m) aboveground in Exposure C:
Columbia River Gorge detail.

Figure 2-22. 100-year MRI radial ice thickness (in.) from freezing rain with
concurrent gust wind speeds (mph) at 33 ft (10 m) aboveground in Exposure C:
Lake Superior detail.
Weather-Related Loads 63

Figure 2-23. 100-year MRI radial ice thickness (in.) from freezing rain with
con­current gust wind speeds (mph) at 33 ft (10 m) aboveground in Exposure C:
Alaska.

thicknesses. Using an ice density ρi = 56 pcf for bubble-free glaze ice, the
linear ice load on a wire is calculated as

Wi = Qi(d + tz)tz (2-18)

where
Wi = Weight of glaze ice (lb/ft customary units, N/m metric units).
Qi = Constant to convert ice thickness to weight (1.2435 customary
units, 0.0282 metric units),
d = Diameter of bare wire (in. customary units, mm metric units),
and
tz = Design ice thickness at height z above ground (in. customary
units, mm metric units)
64 Electrical Transmission Line Structural Loading

Ice accretion on a wire can substantially increase its projected area. The
transverse load due to wind acting on ice-covered wires acts concurrently
with the vertical load due to the weight of the ice. The 3-second gust wind
speeds provided in Figures 2-19 through 2-23 should be used with tz to
compute the ice with concurrent wind load using the methodology pre-
sented in Section 2.1. When calculating forces due to wind on ice-covered
wires, the force coefficient is dependent on the shape of ice buildup
(McComber et al. 1982). However, typical force coefficients of ice-covered
wires are not known. Some organizations recommend using force coeffi-
cients other than 1.0 for wires covered with ice (IEC 2003b, ISO Standard
12494 (ISO 1999)).

2.3.4.4 Design Temperatures for Freezing Rain  Temperature is an


important consideration in calculating the tension of wires. The design
temperatures concurrent with the design ice with concurrent wind loads
due to freezing rain are mapped in Figures 2-24 and 2-25. These
temperatures are for ice thicknesses for all MRIs. It is recommended to use

Figure 2-24. Temperatures concurrent with ice thickness attributable to freezing


rain: 48 contiguous states.
Weather-Related Loads 65

Figure 2-25. Temperatures concurrent with ice thickness attributable to freezing


rain: Alaska.

either these values or 32 °F (0 °C), whichever results in the maximum load
effect. A temperature of 32 °F (0 °C) should be used in Hawaii.
The temperatures shown on these maps were determined by tracking
the minimum temperature that occurred with the modeled maximum ice
load in each freezing rain event. The sample of minimum temperatures for
all the events used in the extreme value analysis of ice thickness was then
analyzed to determine the 10th percentile temperature (i.e., the minimum
temperature that is exceeded by 90% of the recorded minimum tempera-
tures). These 10th percentile temperatures are shown on the maps. In areas
where the temperature contours were near the wind or ice thickness con-
tours, they were moved to coincide with, first, the concurrent wind bound-
aries, and second, the ice zone boundaries.

2.3.5 Ice Accretion on Structural Members


2.3.5.1 Vertical Loads  Ice accretion on the structural members themselves
is typically not included in design. For the design of bracing members of
latticed structures and cross-arms, the construction and maintenance loads
recommended in Chapter 3, Section 3.1 will generally impose design
stresses greater than the bending stresses resulting from the vertical weight
of ice-covered members. For vertical supports (e.g., pole shaft or leg angle),
66 Electrical Transmission Line Structural Loading

the additional axial load due to ice on the member does not significantly
add to the member stress.

2.3.5.2 Concurrent Wind Loads  Ice accretion on the structure may


increase the projected area of the structure exposed to wind loading. For
broad-profile structural members (e.g., pole sections), the fractional
increase in overall projected area due to ice from precipitation icing is
small. For angle members, the increased area may be partially offset by a
reduction in the force coefficient due to the streamlining effect of the ice
coating on the bluff angle member. Thus, for transmission line structures,
it is usually not necessary to design for the increase in the projected area
of a structure due to ice buildup on its members from precipitation icing.

2.3.6 Unbalanced Ice Loads


Although the principal design loading combination is for the same ice
thickness and wind speed applied to all spans, unequal ice loading and
wind speed should also be considered in design. Ice thicknesses and con-
current wind speeds may differ from one span to the next, typically when
the exposure of a transmission line changes as it traverses a hill or ridge.
Generally, tangent structures with longer suspension insulator strings will
not experience significant longitudinal conductor loads due to unbalanced
ice loads; however, suspension structures with short insulator strings, and
in particular shield wire attachments with short hardware assemblies, may
transfer most of the imbalance to the structure. The designer is referred to
Section 3.1 and Appendix I of the manual.

2.3.7 Ice Accretion on Aerial Marker Balls or Similar Devices


Ice accretion due to freezing rain on aerial marker balls should be taken
into account during design. The weight of the ice increases the tension in
the wires and the vertical load on their supporting structures. The ice den-
sity, ρi, is the same as that applied to the wire. The marker ball ice load is
determined using the weight of glaze ice formed on the projected surface
of the aerial marker. Since the ice thickness specified for accretion on wires
is calculated using the diameter to perimeter ratio of 1/π to convert ice
accretion on a flat surface to a cylindrical surface, the wire ice thickness is
multiplied by π as indicated in Equation (2-20). The weight of ice on an
aerial marker ball is calculated using Equations (2-19) through (2-21)

Wi = Viρi (2-19)
Vi = πtzAs (2-20)
As = πr2 (2-21)
Weather-Related Loads 67

where
As = Projected area of aerial marker ball [Equation (2-21)],
Vi = Volume of ice accreted on the aerial marker ball
[Equation (2-20)],
tz = Design wire ice thickness,
r = Radius of the aerial marker ball, and
ρi = Ice density.
CHAPTER 3
ADDITIONAL LOAD CONSIDERATIONS

3.0 INTRODUCTION

Transmission line designers should consider loads and potential failure


scenarios from sources other than the weather-related events described in
Chapter 2. This chapter addresses additional structure and wire system
loads resulting from unbalanced wind and ice, broken conductors, con-
struction and maintenance, and other causes. These loads apply to both
new installations and modifications to existing facilities. Site-specific con-
ditions such as landslides, ice flow, frost heave, and flooding are not
addressed but should be considered during design.

3.1 LONGITUDINAL LOADS, LINE SECURITY, AND FAILURE


CONTAINMENT

3.1.1 Longitudinal Loads


In addition to the transverse and vertical climatic loads discussed in
Chapter 2, transmission line structures are subjected to longitudinal load-
ing due to an imbalance in wire tension or a failed component. These
imbalances can originate from a number of sources and may appear when
the wire system is intact or when it has been compromised. When not
addressed in the design of the structure and line, unabated longitudinal
loading can lead to catastrophic or cascading failures.

69
70 Electrical Transmission Line Structural Loading

3.1.2 Unbalanced Loads on Intact Systems


On suspension structures, the tension in the wire in adjacent spans can
differ due to variable wind speeds or unequal ice accumulation. Wind can
vary from one geographical point to the next due to changes in terrain and
elevation, particularly on long spans or in mountainous regions. Unequal
accumulation of ice can also occur for the same reasons as observed by
in-cloud icing. Lines with limited ability to transfer slack, such as those
with shorter suspension insulators, spans with little sag, and binding of
hardware, can limit the ability of the suspension insulators to swing suf-
ficiently to balance tensions within the wire system.
Strain and dead-end structures are used when a change in tension is
warranted due to line design requirements. On these structures, conduc-
tors are terminated on horizontal insulator strings, which apply tension
directly into the structure. Strain and dead-end structures must be designed
to resist these tension imbalances.
Ground wires are typically assumed to be rigidly attached to all struc-
tures, so an unbalanced tension at this location should be considered for
all structures.

3.1.3 Longitudinal Loads due to Non-Intact Wire Systems


A component failure, such as a wire or insulator, or structure collapse
can create severe load imbalances in the wire system capable of causing the
partial or complete failure of the adjacent structures.

3.1.4 Failure Containment and Line Security Loads


Failures due to unbalanced loading and a broken wire system may con-
tinue to propagate along the line for a significant distance, resulting in a
catastrophic failure. These cascading failures of transmission lines cause
significant damage and large economic losses because they may destroy
complete sections of a line, requiring weeks or months of repair (Ostendorp
1997). One method to prevent these cascading failures is to include line
security loads in the design of transmission structures to provide longitu-
dinal strength. This can be accomplished by including an unbalanced ice
case and/or a broken wire load (BWL) case. Typically, only one phase wire
or the shield wire is considered to be broken for each line circuit.
Many structure types, such as rigid square-based latticed towers, longi-
tudinally guyed V, Y, delta, and portal structures, and single-pole struc-
tures, can be designed with increased longitudinal strength for a relatively
small increase in initial cost. The criticality of the line and the detrimental
effects on the electric system should be considered when determining the
magnitude of the load or whether to apply a line security load at all. It may
Additional Load Considerations 71

be prudent to design structures to minimize cascades and help reduce res-


toration time and cost.
Designers of new lines should be aware of the consequences of a single
structure failure and a potential cascade. In addition to or instead of incor-
porating line security loads or when structures with limited longitudinal
strength are used, localized structure hardening through the use of “stop”
or “anti-cascading” structures may be an effective and economical way to
minimize the impact of cascades. Failure containment philosophies differ
by region, transmission line owner, and system operator. The length and
importance of the line, longitudinal strength of the suspension structures,
terrain, restoration time, emergency stocking levels, cost, right-of-way
access, and proximity to highways and railways will influence the length
of containment sections. Recommended distances between failure contain-
ment structures vary, but separations between 2 and 10 miles (3.2 and
16.1 km) are common.
Several accepted analytical methods can be used to estimate longitudi-
nal design loads to prevent cascading failure. These include the Residual
Static Load (RSL) Method, the Bonneville Power Administration (BPA)
method, Broken Wire Load (BWL) method, and the EPRI method. Refer to
Appendix I for a discussion of these methods and other failure contain-
ment design considerations.

3.2 CONSTRUCTION AND MAINTENANCE LOADS

3.2.1 General
Construction and maintenance (C&M) loads are directly related to work
methods. Construction loads are imposed on structures during assembly
and erection, and from installation of ground wires, insulators, conductors,
and line hardware. Maintenance loads are those loads applied to the struc-
tures resulting from scheduled or emergency inspection and/or replace-
ment of all or part of a structure, ground wire, insulator, conductor, or
conductor hardware system. Knowledge of construction and maintenance
work methods is required to develop appropriate structure loading cases.
The load scenarios described in the following sections should be consid-
ered during design.

3.2.2 Structure Erection


Lifting a structure or components may generate greater stresses in the
members of the structure than those induced by the in-service design loads
(Figure 3-1). The load path may also be different and is determined by the
chosen lift points. The designer should anticipate lifting limitations and
provide attachment locations, as applicable, to control the load path.
72 Electrical Transmission Line Structural Loading

(a)

(b)

Figure 3-1. Structure erection of (a) a tubular steel H-frame structure; and (b) a
latticed steel guyed tower.
Additional Load Considerations 73

3.2.3 Loads Due to Wire Installation


Ground wires and conductors should be installed in accordance with
the recommendations of IEEE Standard 524, “IEEE Guide to the Installation
of Overhead Transmission Line Conductors.” Ground wire and conductor
stringing loads may be larger than the anticipated maximum intact design
loads and are calculated from the stringing geometry and length of wire
pull. Recommended minimum loads and load factors for installing ground
wires and conductors include the following:
• For transverse and longitudinal components of wire tension, use a
tension based on initial wire conditions at the lowest temperature
that is expected to occur during stringing operations. Apply a
minimum load factor of 1.5.
• For transverse wind loads, use a 3 psf (0.144 kPa) wind [35 mph
(15.6 m/s)], the lowest temperature anticipated, and the maximum
design wind span with a minimum load factor of 1.5.
• For dead-end conditions with pulling or tensioning equipment at
ground level, use the vertical component of the pulling line, the
maximum single vertical span, and a minimum load factor of 1.5.
• For intact wire conditions (ahead and back spans are attached to the
structure), use the maximum design weight span and a minimum
load factor of 2.0.
Sections 3.2.3.1 through 3.2.3.6 provide additional detail for specific wire
installation loads that should be considered during the design phase.

3.2.3.1 Wire Tension Loads at Snub Structure  At the ends of a wire pull,
the wire passes over the stringing blocks and then downward to the pulling
or tensioning equipment (Figure 3-2). A pulling line slope of at least three
horizontal to one vertical is typically considered good practice. Because of
this 3H:1V slope, the wire tension produces a vertical load at the location
where the stringing blocks are mounted to the structure. A transverse load
component may develop where the stringing blocks attach to the structure
depending on the location of the pulling equipment. The transverse load
component is a function of the angles made by the wire entering and
leaving the stringing blocks due to the horizontal alignment of the
tensioning equipment (Figure 3-3).

3.2.3.2 Wire Stringing Loads  At structures in the middle of a wire pull,


the wire passes over the stringing blocks. The transverse load component
is a function of the angles made by the wire entering and leaving the
stringing blocks due to the horizontal alignment of the transmission line.
The vertical load component is a function of the conductor weight and the
weight span (Figure 3-4).
74 Electrical Transmission Line Structural Loading

Figure 3-2. Wire stringing operations and vertical and transverse load at snub
structure.

Figure 3-3. Vertical wire tension and wire weight components of load at snub
structure.
Additional Load Considerations 75

Figure 3-4. Intact stringing load conditions.

3.2.3.3 Bound Stringing Block Loads  During tension stringing opera­


tions, the running board may jam in the stringing block. This can damage
the structure when there is inadequate control or time to stop the pulling
operation. Although a few utilities have designed suspension structures to
resist such possible loads, a more practical solution is to control the
stringing operation in accordance with IEEE Standard 524. See Figure 3-5
for a photo of wire stringing operations showing a running board near a
stringing block.

Figure 3-5. Wire stringing operations.


76 Electrical Transmission Line Structural Loading

3.2.3.4 Temporary Wire Attachment Loads  As wire stringing progresses,


the sequence of wire tensioning may require wire tensions to be transferred
from tensioning equipment or temporary anchors to the structure. This
could result in a configuration with some phases or subconductors installed
and temporarily attached to a structure while other wires or phases are not
in place or are being installed. Differential vertical loads and sagging
tension loads should be incorporated in the design of the structure.

3.2.3.5 Vertical Load Increase in Hilly Terrain  When pulling up a slope,


the initial wire tension will increase by the unit weight of the conductor
times the elevation change while the wire is in the blocks and before
clipping-in and offsets are applied. This can severely increase the vertical
load on uphill structures and must be considered in the design of the
structure.

3.2.3.6 Temporary Guy Wire Loads  Guy wires may be used to provide
temporary support for suspension structures, crossarms, or other supports
during stringing operations to control deflection and the load path. These
temporary guy wires increase vertical loads on the structures. The design
capacity of the guy wire system, structure, and structure components
should exceed the anticipated temporary dead-end connection loads.

3.2.4 Maintenance Loads


Maintenance loads act on the structures as a result of scheduled or emer-
gency inspection and/or replacement of all or part of a structure, ground
wire, insulator, conductor, and conductor hardware system. An appropri-
ate load factor should be applied when designing for maintenance loads.
A load factor of 2.0 is generally recommended. Structure maintenance
loads consist of the effects of workers on the structure and of load effects
on adjacent structures due to temporary modifications, such as guying, to
permit the repair or replacement of the structure being maintained. Refer
to Section 3.3 for worker access and fall protection loads.
Common maintenance performed on a transmission line includes
adjusting or replacing ground wires, conductors, insulators, and hardware.
At times, it is necessary to remove the wires from their supports and lower
them to the ground or transfer them to a temporary structure or temporary
alternate location on the structure being maintained. The loads induced
during these operations could exceed the original design loading of the
structure or the adjacent structures.
Prior to lowering wires at one or more structures, the load effects should
be considered. With level spans, the lowering of wires to or near the ground
at one structure will increase the vertical loads at the adjacent structures
and cause the insulators to swing toward the lowered line as the line ten-
sion attempts to equalize. This maintenance operation can create combined
Additional Load Considerations 77

vertical and longitudinal loads, potentially overstressing the adjacent


structures.

3.3 WORKER ACCESS AND FALL PROTECTION LOADS

During erection and maintenance, some structure members are loaded


in flexure by the vertical weight of the workers. This access load should be
applied only to members on which a worker is anticipated to climb or
work. Normally this load is treated as an independent vertical load not less
than 250 lb (1.1 kN), the approximate weight of a lineman with tools. A load
factor of at least 1.5 is recommended.
Climbing devices used exclusively to support personnel, such as step
bolts and ladders, as well as anchorages used for fall protection, may have
additional strength requirements as stated in standards such as Occupa-
tional Safety and Health Administration (OSHA), ANSI C2 National Electric
Safety Code, ANSI/ASSE Z359.1 Fall Protection Code, and IEEE 1307 Standard
for Fall Protection for Utility Work.
Fall protection loads are created when workers, attached to an anchorage,
fall from an elevated position. An anchorage is a secure point of attachment
to which the fall protection system is connected (ANSI C2). The fall protection
system must meet all OSHA 1910.269, ANSI C2, and other legislative require-
ments as applicable. In addition, IEEE 1307 provides guidance regarding
loads and criteria for anchorages and step bolts. The design of the fall arrest
system should be coordinated with operation and maintenance personnel.

3.4 WIND-INDUCED STRUCTURE VIBRATION

Wind-induced vibration can occur with or without the installation of


conductors and shield wires and typically occurs in low-turbulence wind
conditions. This is due to the fact that smoother wind flow leads to a more
organized shedding of vortices from structural members. The frequency at
which vortex shedding occurs for a component is proportional to wind
velocity. When the wind velocity is such that the frequency of vortex shed-
ding matches or is very close to the natural frequency of a component or
system, a condition of harmonic resonance can occur. This mechanism
tends to generate a large number of loading cycles and may result in fatigue
issues or failure. The vortex shedding phenomenon is directly related to
the shape of the member and can often be controlled by altering the profile
of the member as seen by the wind. The amplitude of vibration at harmonic
resonance is directly related to the amount of structural damping and can
be mitigated by the addition of damping mechanisms. Appendix E con-
tains additional information concerning wind-induced structure and arm
vibration and oscillation of structures and members.
78 Electrical Transmission Line Structural Loading

Transmission line structures can be subjected to dynamic forces caused


by wind-induced vibration and conductor motion. These forces have the
potential to initiate vibration in both the complete structure and individual
members. In lieu of a comprehensive engineering analysis considering
local weather conditions and individual structure member characteristics,
the effects of structure vibration may be reduced by one or more of the
following methods: hardware assembly design, structure design, member
detailing, connection detailing, wire vibration dampers, and erection
methodology.
When conductors or ground wires are not installed immediately after
the tubular steel pole arms and crossarms are installed, owners should
install mechanisms to minimize potential wind-induced oscillation. Exam-
ples of these mechanisms include the installation of internal or external
damping devices, internal cables, weights, temporary tiebacks to a fixed
point, and insulator assemblies with travelers (ASCE 2011). The structure
designer should be consulted to determine what measures, if any, should
be used for each specific circumstance.

3.5 WIRE GALLOPING LOAD CONSIDERATIONS

Galloping (high-amplitude, low-frequency wire motion) is a dynamic


load that can occur given the right aerodynamic conditions of
low-turbulence wind blowing across wires with or without a relatively
light ice coating (Figure 3-6). Galloping events are unpredictable and can
occur in one or several phases or spans. Factors influencing the onset of
galloping include wire orientation, diameter, shape, weight, frequency,
damping, and span length. However, galloping is more prevalent in flat
terrain when wind speeds are between 10 and 20 mph (16.1 and 32.2 kmph),
with wind blowing fairly normal to the wires, with an uneven ice coating.
Because of this, random galloping occurrences may be more frequent than
heavy ice loading events. Therefore, the effects of galloping should be con-
sidered in the design of the line when these conditions exist.
The vertical amplitude of the galloping wire can reach or exceed the
wire sag, although most galloping amplitudes are less than 1 m (Den Har-
tog 1932, Davison et al. 1961, EPRI 1979, Havard and Pohlman 1980, Raw-
lins 1981).
The effects of galloping may cause electrical and structural problems
such as
• Flashovers or clashing of wires that lead to temporary or perma-
nent outages due to the reduction of spacing between phases or a
phase and a ground wire (Farr 1980, REA 1980);
Additional Load Considerations 79

Figure 3-6. Conductor galloping.

• Permanent elongation of conductors and ground wires caused by


dynamic wire tensions in the inelastic range. This leads to addi-
tional sag which may infringe on allowable ground clearance (Anjo
et al. 1974, Richardson 1986);
• Excessive wear, fatiguing, and failure of ground wires, conductors,
and associated hardware and insulators of suspension and dead-
end assemblies (EPRI 1979); and
• The collapse of structural components and systems (Baenziger et al.
1993a, b; White 1979).

3.5.1 Wire Galloping Loads


Galloping wires can produce large vertical and longitudinal loads at
supports. Theoretical studies indicate that tensions at dead-ends can vary
by ±60% and the vertical loads at support points can vary by ±30%, with
the magnitude depending on many factors (Brokenshire 1979, Gibbon 1984,
Richardson 1986, CIGRÉ 2007). Measurements on actual galloping lines
have found tension changes at dead-ends, cycling between 80% to 140% of
the static tension (Anjo et al. 1974). It should be noted that cycling of
80 Electrical Transmission Line Structural Loading

vertical loads at support points, which were measured to be of the same


magnitude as the tension changes, may not be visibly evident if the support
point is rigid (Anjo et al. 1974). These vertical and tension loads may cause
structure or structure element failure, such as ground wire masts (White
1979).

3.5.2 Galloping Mitigation


Several methods of galloping mitigation have been implemented. For
new line construction, utility experience has shown that twisted pair con-
ductors can reduce the occurrence of galloping. For existing lines with
round conductors, galloping mitigation measures include detuning pen-
dulums, interphase spacers, airflow spoilers, dampers, and rotational
weights. These and other alternative measures and devices have been
evaluated in field investigations (EPRI 1979, Havard and Pohlman 1980,
Havard et al. 1982, Nigol and Havard 1978, Pohlman and Rawlins 1979,
Whapam 1982). Experiences with mitigation devices indicate varying
degrees of success.
Although increasing the vertical and horizontal spacing between wires
may eliminate flashovers or clashing, this solution may require longer
arms, taller and larger structures, and larger foundations (Boddy and Rice
2009). Additionally, increased wire spacing will not eliminate the other
potential structural problems associated with galloping.

3.6 EARTHQUAKE LOADS

Transmission structures typically need not be designed for ground-


induced vibrations caused by earthquake motion. Historically, transmis-
sion structures have performed well in seismic events as documented in
industry publications, such as ASCE Technical Council on Lifeline Earth-
quake Engineering (TCLEE) and Earthquake Engineering Research Insti-
tute (EERI) reconnaissance reports. Decades of experience with lines of all
configurations support the fact, that few, if any, failures from inertial loads
are seen on transmission structures after an earthquake event. Computer
modeling of both latticed towers and tubular steel structures has shown
that the structure loadings caused by extreme wind, ice, and unbalanced
wire tension loads exceed the loads caused by earthquake events (Riley et
al. 2002). Transmission lines are a complex system of structures, overhead
wires, and insulators. The varying structure types and heights, and cable
span lengths and sags, interconnected with flexible insulators along the
line, result in significantly different lower natural frequencies between the
structures and wires. The advantage of this system is the low relative mass
of the structures and the ability of the cable system to dissipate dynamic
Additional Load Considerations 81

energy. Structure failures during a seismic event have been caused by geo-
technical effects, such as landslides, liquefaction, and lateral spreading.
Thus, the inclusion of seismic inertial loads typically will not control the
design of a transmission structure. The traditional extreme loads, as pro-
vided in this manual, are adequate to obtain the structural capacity to miti-
gate the effects of earthquake inertial loads. The following sections provide
guidance when seismic effects are considered in the design of a transmis-
sion line.

3.6.1 Seismic Hazards


Transmission lines are long continuous structural systems that may
encounter a variety of seismic hazards along potential or established routes.
While the design of transmission structures will typically not be controlled
by seismic loads, structural damage could occur as a result of secondary
ground or terrain effects around a structure. When designing a transmission
line in seismic regions, the following hazards should be considered by
designers to minimize potential transmission structure damage:
• Induced seismic loads on structures subject to strong ground
motion: The transmission structure is part of the line system, which
consists of the structures and supporting wires interconnected by
the insulators. The complexity of this system can vary with struc-
ture types, heights, leg configurations, span lengths, sags, wire
configurations, insulator arrangements, and terrain. To understand
the seismic inertial load on a transmission structure, these param-
eters need be included in a structural finite-element model of the
transmission line system to obtain a representative inertial load.
This type of detailed analysis is typically not performed because of
the lack of structure failures caused by inertia loads during earth-
quake events. For new unique, nontraditional, structure configura-
tions and/or materials, the need for earthquake inertial loads
should be considered.
• Ground rupture and ground movement: Structures built across an
earthquake fault zone could be damaged as a result of permanent
ground displacement due to the fault. Adequate conductor sag and
slack, insulator type, and configuration should be considered to
accommodate large horizontal displacements when siting a trans-
mission line that crosses a fault zone, particularly at a lateral strike/
slip fault.
• Liquefaction: Seismic shaking can result in ground subsidence
(settlement and lateral spreading) under certain soil and ground-
water conditions. Some soils can liquefy during an earthquake and
cause structures that are founded on such materials to experience
82 Electrical Transmission Line Structural Loading

differential foundation settlement and/or large horizontal move-


ment resulting in structural damage or failure.
• Landslides: Seismic shaking can induce landslides which could
either undermine a structure’s foundation or cause slide debris
such as rock falls to impact a structure. The terrain and geology
along the transmission line route should be studied to determine
where these hazards may occur.

3.6.2 Siting and Geotechnical Assessment


Seismic design relative to a transmission line design is normally limited
to a geotechnical earthquake hazard assessment along the line route. Geo-
technical work for a specific project should include an evaluation of areas
susceptible to liquefaction and landslides as well as likely fault rupture
zones. Where seismic activity along a transmission line route is a potential,
a qualified geologist or geotechnical engineer familiar with the seismicity
of the area should be part of the project routing review to provide input for
routing decisions.

3.7 SUMMARY OF ADDITIONAL LOAD CONSIDERATIONS

Table 3-1 summarizes the recommended load factors that are referenced
within this chapter and associated appendixes.
Table 3-1.  Summary of Additional Load Considerations

Recommended load factors


Recommended load case Section Weather conditions Longitudinal Transverse Vertical
Line security loads 3.1.4 See Appendix I
RSL method
EPRI method
BPA method
BWL method
Structure erection loads 3.2.2 — 1.5 1.5 1.5
Wire installation loads
Wire tension loads at snub structure 3.2.3.1 Lowest anticipated ambient tempera- 1.5 1.5 1.5
ture at time of installation; wind = 3
Intact wire loads at tangent structure 3.2.3.2 1.5 1.5 2.0
psf
Bound block loads 3.2.3.2 2.0 2.0 2.0
Maintenance loads 3.2.4 Lowest anticipated ambient tempera- 2.0 2.0 2.0
ture at time of installation; wind = 3
psf
Worker access and fall protection loads 3.3 — 1.5 1.5 1.5
Wind-induced structure vibration 3.4 See Appendix E
Wire galloping 3.5 — — — —
Earthquake loads 3.6 — — — —
CHAPTER 4
WIRE SYSTEM

4.0 INTRODUCTION

A wire system includes conductors, overhead shield wires, and other


cables connected to the structure. Structural wires, such as guy wires, are
not considered part of the wire system. It is necessary to understand the
tension in the wire system as well as the longitudinal, vertical, and trans-
verse loads these wires impose on the structure attachment points. Wire
tensions vary with
• Electrical loading: current and resistance;
• Weather loading: ice, wind, and temperature;
• Wire condition: tensions at initial, final after creep, and final after
load;
• Structure support flexibility: structure type and insulator
configuration;
• Construction and maintenance: stringing and erection techniques;
and
• Span characteristics: span lengths and elevation changes.
The effects listed above alter the temperature, length or support point,
and therefore tension, of each wire. Wire tensions temporarily decrease as
wire temperature increases and thermally expands (elastically lengthens)
the wire. Conversely, tensions increase with decreased temperatures. Simi-
larly, tensions decrease as wires creep (plastically lengthen) with load and
time. External loading of ice and wind will also increase wire tension.
These and other loading issues will cause any given wire to experience a
range of tensions. The wire tensions must be contained within limits to

85
86 Electrical Transmission Line Structural Loading

ensure the viability and performance of the wires and of other components
of the wire systems.
The wire tensions directly affect
• Longitudinal loads applied to the strain and dead-end structures,
• Transverse loads at all line angles,
• Factored longitudinal design loading or Residual Static Load (RSL)
to resist cascades (See Chapter 3),
• Vertical loads at structures with a vertical angle (VA) (See Figure 4-1),
and
• Foundation, guy wires, and anchors.
Loads per unit length of conductor or ground wire have been discussed
in Chapter 2. This chapter discusses the manner in which the wire system
is affected and the assumptions that may be used for determining the loads
at the structure attachment points.

4.1 TENSION SECTION

The wire system is normally broken down into tension sections. A ten-
sion section is a portion of conductor or ground wire strung between
dead-end points, such as between Points A and E in Figure 4-1. Wire ten-
sion only affects structures where a horizontal or vertical angle occurs in
the tension section and where wires terminate. If the wires at the support-
ing structures do not have a horizontal angle (Points B and D in Figure 4-1),
then the transverse loads are not affected by wire tensions. For structures
with a vertical angle (Location B in Figure 4-1), the vertical and

Figure 4-1. (a) Plan; and (b) profile of tension section.


Wire System 87

longitudinal load could be affected by the wire tensions depending on the


geometry of the spans.
Intermediate wire attachment points B, C, and D between the dead-end
points are presumed to have some longitudinal flexibility. This flexibility,
or freedom to move, equalizes the horizontal components of tension for
various loading events. Longitudinal flexibility in the wire system is pro-
vided by two possible sources: insulator configuration and structure type.
At Points B, C, and D, suspension insulators allow more wire movement
than post insulators or overhead ground wire clamps. Poles or H-frames
exhibit more longitudinal flexibility than more rigid lattice towers or guyed
structures. The importance and complexity of the wire system longitudinal
flexibility is discussed in Section 4.4.1.

4.2 WIRE CONDITION

Wires, especially conductors, are subject to permanent elongations over


their lifetime in service. They are said to be in their “initial” condition if
they are new wires and within a few hours of installation. The wires are in
their “final after creep” condition if they have been in service for several
years and have experienced permanent elongation over time under normal
operation. The creep process slows considerably with time. Estimates of
creep are usually based on a 10-year period. Therefore, the wire will spend
most of its life at a condition close to what is commonly called “final after
creep.” The majority of creep elongation occurs in the initial one or two
years following the conductor installation.
Assumptions regarding wire elongation and “initial” or “final” condi-
tions are based on wires of aluminum, steel, and/or copper, commonly
used for electrical transmission lines. In recent years, high temperature–
low sag conductors have been developed, which have different material
characteristics.
The wires are in their “final after load” condition if they have been per-
manently elongated by a high tension due to the maximum design load,
such as an extreme load from ice, wind, or a combination of both as
described in Chapter 2. Permanent elongation from heavy load can be
determined more precisely than elongation from creep if future load func-
tion or load history can be accurately predicted. However, since the selec-
tion of a weather-related load (used in the calculation of wire tension) is a
design assumption (future estimation of applied load), the permanent elon-
gation calculated is also an estimation. The magnitude of the “final after
load” cases, selected for design, is intended to be an upper limit and should
be evaluated using computational methods available in typical design soft-
ware. Since “final after load” depends on extreme weather cases such as
extreme wind or heavy ice, it is possible that this load condition may never
occur over the life of the line.
88 Electrical Transmission Line Structural Loading

The total permanent elongation of the conductor under the impact of


creep or load will depend on the time loading history of the conductors.
For example, if a severe weather load occurs very early in the life of the
line, the permanent load elongation will cause the everyday conductor
tension to decrease.
An accepted practice in the industry is to disassociate creep and load
permanent elongation and assume them to be independent and not addi-
tive. Compared to their initial values, tensions are lower for “final after
creep” and “final after load” conditions because the wires have perma-
nently elongated. A situation where a “final after creep” tension is lower
than the corresponding “final after load” tension is sometimes referred to
as “creep is a factor” or “creep controls.” For certain design aspects such
as aeolian vibration, it is best to consider both wire states. An alternate
method has been offered by CIGRÉ Task Force B2.11.04 (2004) for tension-
ing of wires with respect to aeolian vibration.
In addition to sag-and-tension related wire conditions such as initial and
final, conductors, ground wires, and other components of the wire subsys-
tem can deteriorate. Some electrical, environmental, and other physical
impacts on wire condition are
• Lightning strikes,
• Wire motion fatigue,
• Corrosion of steel strands,
• High electrical loading,
• Impact damage,
• Electrical flashovers, and
• Hardware fatigue.

4.3 WIRE TENSION LIMITS

It is critically important to protect the wire system because failure of a


wire, or any components in series with the wire, will impose longitudinal
loads on the structure system that may exceed design loads causing failure.
Wires are normally sagged to perform within certain design limits. Limits
on everyday tensions or on catenary constants under initial and/or final
conditions are normally given to avoid or to minimize the potential for
wind vibration damage.
For tension limits, legislated requirements such as NESC or IEC 60826
along with wire manufacturer recommendations should be followed.
Many utilities specify more stringent tension limits. The everyday wire
condition should be specified as “final after creep” and not “final after
load” since the “final after creep” generally has a higher tension magnitude
for the reasons discussed in Section 4.2. For mitigation of wire vibration,
the tension associated with “final after creep” is typically considered a
Wire System 89

prudent design practice as the wire will be in that tension condition for a
majority of its service life.
Designing using “final after load” may result in a wire enduring higher
tension for its service period until the heavy load occurs. This can cause
unforeseen vibration problems due to a higher tension. Wire vibration miti-
gation devices should be installed with appropriate consideration of wire
tensions, span lengths, and local weather conditions.
When extreme loads do occur on the wire systems, all the components
in series within these systems are highly stressed and must have adequate
remaining strength. Therefore, it is recommended that the maximum ten-
sions caused by the factored extreme loads never exceed 70% to 80% of the
rated tensile strength of the wire.

4.4 CALCULATED WIRE TENSION

4.4.1 The Ruling Span Method


Assuming that the horizontal components of tension, TH, in all the spans
of a tension section are the same when the spans are in relatively flat ter-
rain, then the entire tension section can be represented as a single equiva-
lent “ruling” span (Thayer 1924). The wire horizontal tension, TH, for a
given set of temperatures and external loads may be determined by sub-
jecting the ruling span to those conditions.
The ruling span method implies that the same unit load is applied on
all the spans of the tension section. To make this assumption valid, the
intermediate support points must have sufficient longitudinal flexibility to
allow this equalization of tensions. This method is valid for wire sag and
tension calculations where span lengths are similar. Rugged mountain ter-
rain and sizeable span length differences reduce the accuracy of the ruling
span model. The ruling span is an approximation that has limitations (IEEE
1999). The ruling span is computed as follows:

S13 + S23 + S33 +  + Sn3


Ruling Span = (4-1)
S1 + S2 + S3 +  + Sn

where S1, S2, S3, . . ., Sn is the individual span lengths (horizontal projec-
tions) between dead-end or strain structures.
When correctly applied, the ruling span method [Equation (4-1)] enables
the stringing and sagging-in of a line section (i.e., between dead-ends) of
unequal spans in flat or hilly terrain so that the horizontal tensions in each
span will be equal as designed.
The ruling span method relies on the ability of the wire connection
points to move longitudinally without restriction as a means to equalize
90 Electrical Transmission Line Structural Loading

tensions between spans. Calculated tensions may not be as accurate for


conductor spans supported by rigid post insulators or for ground wires
supported by rigid clamps. They may also not be as accurate if suspension
insulators are not sufficiently free to move in the longitudinal direction
(e.g., at support points with substantial horizontal and vertical line angles).
In such cases, the structural analysis options described in the next subsec-
tions are more appropriate.
The ruling span method is not applicable to the calculation of unbal-
anced longitudinal loads caused by restricted suspension points, uneven
ice on adjacent spans, or other span-specific disturbances.

4.4.2 Structural Analysis of a Single Tension Section


If there is no significant interaction between consecutive tension sections
caused by the longitudinal displacement of their supporting structures,
then the tensions in the various spans of a tension section can be deter-
mined by modeling that tension section as a wire system with appropriate
support conditions. Suspension supports can be modeled as cable elements
or swinging rods. Post insulators can be modeled as small, cantilevered
beams or longitudinal springs with appropriate longitudinal flexibilities.
The model is then analyzed by accepted structural analysis methods that
account for the longitudinal displacements of the wire attachment points
while incorporating the corresponding changes in wire tensions.
This type of analysis is capable of handling unbalanced ice and/or wind
loads and will produce more accurate tension results than the ruling span
method.

4.4.3 Structural Analysis of Multiple Tension Sections


If there is significant interaction between consecutive tension sections
caused by the longitudinal displacement of their supporting structures,
then those tension sections can be analyzed as a single structural system,
which includes all the wires in all the spans as well as detailed structural
models of all supports. This rigorous approach produces a more accurate
analysis but, because of its complexity, is normally only justified in special
situations.

4.4.4 Computational Methods


Modeling of a tension section or multiple tension sections depends on
the wire system and support modeling methods and assumptions since
they operate in concert under various load cases. For 2D models where the
ruling span, actual wind/weight span, attachment displacement, and indi-
vidual conductor line angle may all, or in part, be held constant under
different load cases for computational simplification, the results may be
Wire System 91

divergent and varied. This is because these effects may be too critical to
ignore, especially if the tension sections have significant combinations of
line angles, permanent structure deflections, and/or elevation differentia-
tion. In 3D models that incorporate lateral and longitudinal movements for
conductor, support hardware, and structure, the computational results
reflect a more accurate result than 2D models.

4.5 LOADS AT WIRE ATTACHMENT POINTS

4.5.1 Wire Unit Loads


A length of wire, as received from the manufacturer, possesses certain
physical and electrical properties making it suitable for use on a transmission
line. Electrical properties, such as the area of aluminum, or other conductive
material, allow it to transfer current between one station and another. The
physical and dimensional properties of that aluminum, in connection with
steel or other strength components in the conductor, give resultant diame-
ters, weight, and tensile strength. These basic wire properties of diameter,
weight, and tensile strength are associated with the major transverse, verti-
cal, and longitudinal loads imposed on their supporting structures.
The basic loads applied to wires can be divided into vertical (weight),
transverse (wind), and longitudinal (tension) components (Figure 4-2). Ver-
tical load (applied downward) and horizontal or transverse loads (applied
perpendicular to the line) are calculated on a unit-of-length basis. These
load components can be resolved into a resultant vector load. Longitudinal
load (applied along the line), due to wire tension, is not typically consid-
ered as a unit load.
Bare wire will have the dimensional properties and stress versus strain
characteristics as given by the supplier, yet wire in service is exposed to
varying weather and construction and maintenance loads, which cause the
unit loads to change with environmental conditions and loading history.
Horizontal load is a function of wind pressure as discussed in Chapter 2
applied to the projected area of the wire, which may increase with accu-
mulated ice or snow. Vertical load is a function of the bare wire weight
coupled with that of any ice or snow. The Vertical Unit Load, wv, and Hori-
zontal Unit Load, wh, are calculated using Figure 4-2. These unit wire loads
vary with loading history and weather conditions as described. All signifi-
cant load cases should be considered in structure design.

4.5.2 Using Wind and Weight Spans


Vertical and transverse unit wire loads, wv and wk, are multiplied by the
applicable wind or weight span to obtain the wire loads applied to the
92 Electrical Transmission Line Structural Loading

Figure 4-2. Unit wire loads.

structures. At tangent locations, such as Points B and D in Figure 4-1, the


transverse and vertical structure loads, LT and LV, can be determined as

LT = load factor × wh × wind span (4-2)

LV = load factor × wv × weight span (4-3)

where wind span is the length of wire between midspan points in the
adjacent spans. In traditional (2D) calculations, the wind span was calcu-
lated as one half of the horizontal projections of the adjacent spans as
shown in Figure 4-3. Weight span is the length of wire between the low
points in the adjacent spans. In traditional (2D) calculations, the weight
span was calculated as the horizontal distance between the low points in
the adjacent spans as shown in Figures 4-3 and 4-4.
Three formulas are presented to locate the low point of a span (Winkel-
man 1959): an approximate equation [Equation (4-4)] or more precise cat-
enary equations [Equations (4-5) and (4-6)]. It must be noted that Equation
(4-4) should not be used if the difference in support elevations (B) is greater
than approximately 20% of the span length (S).
Wire System 93

Figure 4-3. Wind and weight spans.

Approximate Formula. The position of the low point of sag, XL, is approxi-
mated by the following formula:

S
XL = × (midspan sag − Bi/4)
2× midspan sag

sagm = midspan sag = (Si × Di)/(8 × C) (4-4)

Di = Si2 + Bi2

where subscript i denotes the span being considered from Figure 4-3.
Calculating the weight span for a particular wire loading requires deter-
mining the equilibrium configuration of the wire for that loading.
These low points can also be determined using the hyperbolic equations
for a catenary shown in Equations (4-5) and (4-6). As terrain difference
increases (slope angle VA becomes large), the calculated arc length of the
wire l [Equation (4-7)] between the low point and the support should be
used for calculating LV using Equation (4-10).
94 Electrical Transmission Line Structural Loading

 
 
 B 
S −1  2
X L = − C × sinh   (4-5)
2  
 S  

 C × sinh   
  2×C  

 
 
 B 
S
XU = + C × sinh−1  2  = S − XL (4-6)
2   S  

 C × sinh  
  2×C  

 X
l = C sinh  (4-7)
 C

C = TH/wr (4-8)

where
C = Catenary constant, which is the ratio of horizontal tension to
the unit wire load,
S = Span length,
B = Difference in elevation of supports,
D = Straight-line distance between the supports,
XL = Distance from low point of sag to lower support,
XU = Distance from low point of sag to upper support,
l = Length of wire from low point to support [X = XL for lower
support and X = XU for upper support in Equation (4-7)],
TH = Horizontal component of tension, and
wr = Unit wire load for the desired load case.

Equations (4-4) to (4-8) assume that there is no blow-out of the spans,


which is only true if wr equals wv (wk = 0).
In such a case, the cables in Figure 4-3 are in a vertical plane. With some
wind, the planes of the cables rotate around the straight line between the
supports in the direction of wr. However, the general aspect of Figure 4-3
remains valid, but with different sags and locations of the low points. These
sags and low points can still be located using Equations (4-4) to (4-8) if one
replaces the catenary constant of Equation (4-8) by TH/wv.
At line angle locations, such as Structure C in Figure 4-1, the horizontal
component of wire tension in the adjacent spans results in transverse load.
Wire System 95

This additional transverse load is calculated in Equation (4-8) and should


be added to the transverse load in Equation (4-2)

LT-Angle = 2 × TH(sin0.5HA) (4-9)

where TH = horizontal component of tension, and HA = horizontal line


angle.
A good approximation of the vertical load, LV, can also be obtained by
Equation (4-10) for inclined spans versus wire arc length calculation.

LV = factored unit vertical wire load


× wind span + TH × 2 × tan[0.5 × VA] (4-10)

where VA = vertical line angle.

4.5.3 Weight Spans on Inclined Spans


The low point of sag for structures with equal wire attachment eleva-
tions occurs at the midpoint of the span. The most simplistic calculation of
weight span occurs when three consecutive structures support the wire at
the same elevation. In this case, the weight span applied to the center struc-
ture equals half the sum of the two adjacent span lengths.
Where wire attachment elevation points are not equal, the low point of
sag will not coincide with the span midpoint. When slope increases,
whether in terrain or in structure height, the low point will move away
from the higher attachment point. The situation is compounded with
changes in tension due to temperature- or weather-related events. With
decreasing temperature, the wire length decreases, the wire tension
increases, and the catenary curve created by the suspended wire becomes
more flat. This also moves the sag low point (Figure 4-4).
In the “Maximum Sag” portion of Figure 4-4, it is clear to see the low
points occurring in each span and between each structure. The weight
spans for Structures B, C, and D are shown as the distance between the
respective low points of sag. The “Minimum Sag” portion, however, shows
that increased wire tension flattens the curve of the wire catenary and can
result in there being no low point (or belly) of sag within the actual span.
Projected low points in the figure are indicated by dotted lines.
Wire tension changes may increase or decrease the weight span of a
given structure depending on terrain. In the case of Structure D, the pro-
jected low points for minimum sag are beyond the structure from both
adjacent spans. Structure D therefore has a negative weight span, meaning
vertical loads through wires are upward rather than downward and are
trying to lift the structure.
96 Electrical Transmission Line Structural Loading

Figure 4-4. Weight span for inclined spans based on varying sags.

Transmission designers must account for this varying weight span and
its effect on structure loading. Structures subjected to a negative weight
span (or uplift) under certain loading conditions should use insulator and
hardware assemblies designed to resist these vertical forces. These upward
vertical loads must also be considered in the structure and foundation
designs.
Conversely, structures must be designed to support increased vertical
loads as the weight span is lengthened when wire tension decreases.
Knowing the range of possible weight spans for a structure allows the
application of the proper range of vertical loads in the structure and foun-
dation designs.
Equations (4-2) and (4-3) can be used to calculate weight load distribu-
tions based on tension and elevation conditions.

4.5.4 Weight Span Change with Blow-Out on Inclined Spans


Extreme transverse winds on inclined spans can result in large horizon-
tal conductor displacement, commonly known as blow-out. This can pro-
duce changes to the weight spans at the supports. The location of the low
point of the sag (or projected low point) can shift dramatically in the span.
When slopes exceed approximately 20%, the weight span at the upper
Wire System 97

support point can approach double that calculated by methods which


ignore the blow-out, and may even exceed the factored construction and
maintenance loads. Furthermore, the reduction of weight at the lower point
may lead to excessive swing of the insulator strings and result in inade-
quate electrical clearances to the structure. Also, the net vertical force may
be an uplift that can collapse the crossarm. Some line design computer
programs include calculation of the shift of weight spans with blow-out of
inclined spans. The methods discussed in Section 4.5.1 provide a means to
determine if this condition exists [Equations (4-5) and (4-6)].
If the span inclination exceeds about 25% and extreme winds are to be
expected, more precise calculations may be considered or else conservative
vertical load values may be applied to the structures at the upper and lower
support points. These calculations can be made analytically (Keselman and
Motlis 1996, 1998) or with a finite-element computer program (Peyrot 1985)
by breaking the spans into short cable elements, with each element
responding individually to a different local wind incidence. It is important
to note that the analysis of this severe blow-out problem is very dependent
on the horizontal angle at which the wind strikes the span and on the verti-
cal angle of approach of the wind. Both of these angles are likely to vary
considerably and randomly in rough or mountainous terrain where steeply
inclined spans are to be found. Deviations from the orthogonal can greatly
increase these distortions of the wire systems and increase or decrease the
expected weight span changes. Thus, if a serious blow-out problem is
anticipated, there is even greater justification for a conservative approach
to the strengths of the upper and lower structures.
Weight spans of structures vary with temperature due to the wire length
changing with temperature. The weight spans of a structure with higher
wire attachment elevations than the adjacent structure will increase as the
wire temperature decreases. This is due to the low point of the sag moving
away from the higher attachment points toward the lower attachment
points (downhill) as the wire contracts due to the temperature change. The
opposite is true at structures that are lower than adjacent structures. There-
fore, in no-ice areas the largest vertical load, which can occur at a higher
structure, is generally caused by the coldest temperature or sometimes by
wind blow-out. Similarly, the coldest temperature or the wind blow-out
can cause uplift and insulator swing problems at lower structures.
In icing areas, the weight under ice should be used to calculate the verti-
cal load with Equation (4-3). For higher towers, it will almost certainly be
found that the iced weight span is substantially less than the cold bare-wire
weight span.
Equations (4-2) and (4-3) can be used to determine design loads on a
new family of structures intended to have transverse and vertical capabili-
ties based on assumed maximum (allowable) wind and weight spans.
98 Electrical Transmission Line Structural Loading

When these structures are spotted, their ability to carry their design loads
at a particular location is simply checked by verifying that the actual (as
spotted) wind and weight spans are less than the allowable values. In icing
areas, the fact that iced weight spans are generally shorter than cold
bare-weight spans can be used to advantage by specifying shorter allow-
able weight spans under ice than under bare cold.
The concept of allowable wind and weight spans is extremely useful
when spotting new lines, especially with families of standardized struc-
tures. However, for the design of custom structures at specific locations,
for the checking of existing lines, or for parametric studies for possible
upgrading or reconductoring, there is no need to be concerned with
approximations in the wind and weight spans approach if the loads are
computed by a structural analysis method that accounts for the 3D behav-
ior of the wire system.

4.5.5 Centerline Horizontal Angle versus Wire Horizontal Angle


When the wire configuration changes either by rolling phases or arm
length changes, the wire angle of one or more phases can vary from the
centerline angle. For design purposes, the centerline angle is often
employed as the wire horizontal angle for loading purposes. For a line
design that utilizes the same structure geometry, this is a valid general
assumption. However, it is important to remember that the loading should
be calculated using the wire angle rather than the centerline angle when
the two differ from one another. A given structure may have several differ-
ent wire angles depending on the structure geometry and wire arrange-
ment. For these structure locations, the assumption of the centerline angle
representing the wire loads may not be sufficiently accurate for the
intended purposes.
Examples of this include
• Vertical to horizontal phase rolling, or phase transpositions;
• Vertical to delta configuration;
• Other structure type changes with different phase geometries; and
• Alignment convergence.
CHAPTER 5
EXAMPLES

5.0 LATTICED SUSPENSION TOWER LOADS

This example shows calculations for wire and structure loads. The loads
are based on the tower shown in Figure 5-1, Tables 5-1 through 5-3, and the
design and wire data listed below. Notation within this chapter has been
kept consistent with other chapters where possible. For convenience, the
coordinate system and variable definitions below only apply to this chap-
ter. Calculation results have been rounded to simplify presentation in the
tables. Nothing is to be inferred from the rounding methods used in this
chapter.
Structure Coordinate System:
• Vertical axis: Local axis that is parallel with the direction of gravita-
tional force.
• Longitudinal axis: Local axis that is parallel with the general
direction of the transmission line conductors and perpendicular to
the vertical axis.
• Transverse axis: Local axis that is perpendicular to the plane
formed by the vertical and longitudinal axes.
Loading nomenclature:
• V = Structure load that is parallel with the vertical axis.
• L = Structure load that is parallel with the longitudinal axis.
• T = Structure load that is parallel with the transverse axis.
• H = Horizontal component of wire tension.

99
100 Electrical Transmission Line Structural Loading

5.0.1 Design Data


Location: Utah
MRI = 100 years
90 mph extreme wind speed
0.25 inch ice, 40 mph concurrent wind
Ruling span = 1250 ft
Wind span =1500 ft
Weight span = 1800 ft
Line angle = 5°
Length of insulator assembly = 6 ft
Weight of insulator assembly = 200 lb
Weight of shield wire assembly = 50 lb
No topographic effects, Kzt = 1.0
Wind pressures on wires for the two loading cases Wind and Wind
at 30° Yaw Angle are calculated based on the wind span. For
calculations of sags and tensions, the use of the ruling span is
appropriate.

Figure 5-1. Suspension tower.


Table 5-1.  Wire Data
954 kcmil 45/7 ACSR 7#8 aluminum clad steel
rail conductor shield wire (d = 0.385 in.,
(d = 1.165 in., w = 1.075 lb/ft) w = 0.262 lb/ft)
Temperature Ice Wind Initial
Loading case (°F) (in.) (psf) Initial sag (ft) Initial tension (lb) Initial sag (ft) tension (lb)
Wind 60 0 16.1 43.5 8,530 38.8 2,917
Wind at 30° 60 0 12.1 42.1 7,396 37.0 2,472
Wind and ice 15 0.273 3.17 39.3 8,092 36.3 2,929
C&M 15 0 3 36.5 5,971 31.0 1,757
FC 30 0 0 37.4 5,622 31.4 1,628
No wind 60 0 0 39.6 5,305 33.0 1,550
Note: C&M = construction and maintenance, FC = failure containment load.
Table 5-2.  Load Summary (kips)
Transverse wind Longitudinal
Shield wire Conductor on structure wind on structure
Loading case V T L V T L FT1 FT2 FT3 FL1 FL2 FL3
Wind 0.5 1.0 — 2.1 3.1 — 1.2 2.0 3.9 — — —
Wind at 30° 0.5 0.8 — 2.1 2.4 — 1.5 2.0 4.0 0.9 1.2 2.3
Wind and ice 0.9 0.6 — 3.0 1.4 — 0.2 0.4 0.8 — — —
C&M 1.3 0.4 — 4.7 1.4 — 0.3 0.5 0.9 — — —
FC (broken wire) 0.3 0.1 1.6 1.2 0.2 3.9 — — — — — —
FC (intact wire) 0.5 0.1 — 2.1 0.5 — — — — — — —
Note: C&M, construction and maintenance, FC = failure containment load, V = vertical load, T = transverse
load, and L = longitudinal load.
102 Electrical Transmission Line Structural Loading

Table 5-3.  Weight Span Summary

Upper tower Lower tower


Weight Difference Weight Difference
Wire C span (ft) (%) span (ft) (%)
Shield Traditional 1652 +23% 424 −136%
wire
Section 5.1 2140 — 180 —
Conductor Traditional 1608 +15 446 −45%
Section 5.1 1884 — 308 —
Note: C-catenary constant or parameter of the catenary curve.

5.0.2 Extreme Wind (Chapter 2, Section 2.1)


For purposes of this example, the wind is assumed to be normal to the
ahead span, back span, and to the structure. Note that the effective height
of the wire system in this example was conservatively taken as the arm
attachment height.
The structure is located on the perpendicular bisector of the line angle.
From the wind map (Figure 2-1 in Chapter 2), V100 = 90 mph, Exposure
Category C.

5.0.2.1 Wind on Wires


3 (74) + 2 (89)
Average wire height zh = = 80 ft from Section 2.1.4.3
5

2
2
 z a  80 9.5

K z = 2.01 h  = 2.01  = 1.21 (2-3)
 z g   900 

1 1
 33 6  33 6
I z = cexp   = 0.2   = 0.17 (2-6)
 z   80 
h

1 1
Bw = = = 0.394 (2-8)
0.8S 0.8 (1500)
1+ 1+
Ls 220
Examples 103

 1 + 4.6 I z Bw   1 + 4.6 (0.17 )(0.394)


Gw =  =  = 0.64 (2-5)
 1 + 6.1I z   1 + 6.1(0.17 ) 

F 2
Wind pressure = QK z K zt (V100 ) GC f (2-1a)
A

2
= (0.00256)(1.21)(1.0)(90) (0.64)(1.0)

= 16.1 psf

5.0.2.2 Shield Wire Loads


V = vertical = w × weight span + hardware weight

= 0.262(1800) + 50 = 522 lb = 0.5 kips


T = transverse = wind pressure × A + 2(wire tension )sin
2
 0.385   5
= 16.1 
 (1500) + (2)(2917 ) sin   = 1029 lb = 1.0 kips
 12   2 

5.0.2.3 Conductor Loads


V = vertical = w × weight span + hardware weight

= 1.075(1800) + 200 = 2135 lb = 2.1 kips



T = transverse = wind pressure × A + 2(wire tension )sin
2
 1.165   5
= 16.1 1500) + (2)(8530) sin   = 3089 lb = 3.1 kips
 12 (  2 

5.0.2.4 Wind on Structure

2 (89)
Two-thirds of the structure height zh = = 59.3 ft from
3
Section 2.1.4.3

2
2
 z a  59.3 9.5

K z = 2.01 h  = 2.01 = 1.13 (2-3)
 z g   900 
104 Electrical Transmission Line Structural Loading

1 1
 33 6  33 6
I z = cexp   = 0.2  = 0.18 (2-6)
 z   59.3 
h

1 1
Bt = = = 0.932 (2-7)
0.56 zh 0.56 (59.3)
1+ 1+
Ls 220

 1 + 4.6 I z Bt   1 + 4.6 (0.18)(0.932)


Gt =  =  = 0.844 (2-4)
 1 + 6.1I z   1 + 6.1(0.18) 

F 2
Wind pressure = QK z K zt (V100 ) GC f (2-1a)
A

= (0.00256)(1.13)(1.0)(90)2(0.844)(1.0)

=19.8 psf
Figure 5-1 shows tower areas and solidity ratios.

2
Transverse wind loads C f = 4.0Φ − 5.9Φ + 4.0

For ΦT 1 = 0.69, C fT 1 = 4.0(0.69) − 5.9 (0.69) + 4.0 = 1.83 , and AT 1 = 34 ft


2 2

For ΦT 2 = 0.17 , C fT 2 = 4.0(0.17 ) − 5.9 (0.17 ) + 4.0 = 3.11, and AT 2 = 33 ft


2 2

2
For ΦT 3 = 0.15, C fT 3 = 4.0 (0.15) − 5.9 (0.15) + 4.0 = 3.21, and AT 3 = 62 ft
2

where force coefficient equations are from Table 2-4

FT 1 = 19.8 (1.83)(34) = 1232 lb = 1.2 kips

FT 2 = 19.8 (3.11)(33) = 2032 lb = 2.0 kips

FT 3 = 19.8 (3.21)(62) = 3941 lb = 3.9 kips

5.0.3 Wind at 30°: Extreme Wind at 30° Yaw Angle (Chapter 2, Section 2.1)
Wind is at a 30° yaw angle.
Examples 105

5.0.3.1 Wind on Wires  From the Wind load case, the wind pressure
normal to the wires equals 16.1 psf

Wind pressure = 16.1cos2 (ψ) = 16.1cos2(30) = 12.1 psf (2-12)

Shield Wire Loads


V = 0.262(1800) + 50 = 522 lb = 0.5 kips
 0.385   5
T = 12.1 1500) + (2)(2472) sin   = 798 lb = 0.8 kips
 12 (  2 

Conductor Loads
V = 1.075 (1800) + 200 = 2135 lb = 2.1 kips

 1.165   5
T = 12.1 1500) + (2)(7396) sin   = 2407 lb = 2.4 kips
 12 (  2 

Wind on Structure
From the Wind load case, the structure wind pressure equals 19.8 psf.
Transverse force coefficients and areas are provided in the Wind loading
case

2
Wind force = QK z K ztVMRI Gt (1 + 0.2sin 2 (2 Ψ ))(C ft Amt cos 2 ¨ + C fl Aml sin 2 Ψ )

(2-14a)

C f = 4.0Φ 2 − 5.9Φ + 4.0 from Table 2-4

For Φ L1 = 0.26 , C fL1 = 4.0(0.26) − 5.9 (0.26) + 4.0 = 2.74 , and AL1 = 43 ft
2 2

For Φ L 2 = 0.24 , C fL 2 = 4.0(0.24) − 5.9 (0.24) + 4.0 = 2.81, and AL 2 = 37 ft


2 2

2
For Φ L 3 = 0.16 , C fL 3 = 4.0 (0.16) − 5.9 (0.16) + 4.0 = 3.16 , and AL 3 = 67 ft
2

where force coefficient equations are from Table 2-4.

Transverse Wind Loads


Transverse wind force = 19.8 (1 + 0.2sin 2 (2 Ψ ))×

(C ft Amt cos 2 Ψ + C fl Aml sin 2 Ψ )(cosΨ ) (2-14a and 2-14b)


106 Electrical Transmission Line Structural Loading

FT 1 = 19.8 (1 + 0.2sin 2 (2 (30)))((1.83)(34) cos 2 (30) + (2.74)( 43) sin 2 (30))(cos (30)) = 1501 lb = 1.5 kips

FT 2 = 19.8 (1 + 0.2sin 2 (2 (30)))((3.11)(33) cos 2 (30) + (2.81)(37 ) sin 2 (30))(cos (30)) = 2030 lb = 2.0 kips

FT 3 = 19.8 (1 + 0.2sin 2 (2 (30)))((3.21)(62) cos 2 (30) + (3.16)(67 ) sin 2 (30))(cos (30)) = 3987 lb = 4.0 kips

Longitudinal Wind Loads


Longitudinal wind force =
19.8 (1 + 0.2sin 2 (2 Ψ ))(C ft Amt cos 2 Ψ + C fl Aml sin 2 Ψ )(sinΨ )

(2-14a and 2-14c)

FL1 = 19.8 (1 + 0.2sin 2 (2 (30)))((1.83)(34) cos 2 (30) + (2.74)( 43) sin 2 (30))(sin (30)) = 867 = 0.9 kips

FL2 = 19.8 (1 + 0.2sin 2 (2 (30)))((3.11)(33) cos 2 (30) + (2.81)(37 ) sin 2 (30))(sin (30)) = 1172 = 1.2 kips

FL3 = 19.8 (1 + 0.2sin 2 (2 (30)))((3.21)(62) cos 2 (30) + (3.16)(67 ) sin 2 (30))(sin (30)) = 2302 = 2.3 kips

5.0.4 Extreme Radial Glaze Ice with Wind (Chapter 2, Section 2.3)
Wind on Wires
From the Wind load case, Kz equals 1.21 and Gw equals 0.64.
From the wind and ice map [Figure 2-19(a)], t100 equals 0.25 in. and VI
equals 40 mph.
2
Wind pressure = QK z K zt (VI ) Gw C f (2-1a)

= 0.00256(1.21)(1.0)(40)2(0.64)(1.0)

= 3.17 psf

0.10 0.10
 z  80 
tz = t100   = (0.25)  = 0.273 in. (2-17a)
 33   33 

Shield Wire Loads


Wi = 1.24 (d + tz )(tz ) (2-18)

lb
= 1.24 (0.385 + 0.273)(0.273) = 0.223
ft
di = 2(0.273) + 0.385 = 0.931 in.
Examples 107

V = (0.262 + 0.223)(1800) + 50 = 923 lb = 0.9 kips

 0.931   5
T = 3.17  1500) + (2)(2929) sin   = 624 lb = 0.6 kips
 12 (  2 

Conductor Loads
Wi = 1.24 (d + tz )(tz ) (2-18)

lb
= 1.24 (1.165 + 0.273)(0.273) = 0.487
ft
di = 2(0.273) + 1.165 = 1.711 in.

V = (1.075 + 0.487 )(1800) + 200 = 3012 lb = 3.0 kips

 1.711   5
T = 3.17  1500) + (2)(8092) sin   = 1384 lb = 1.4 kips
 12 (  2 

Wind on Structure
From the Wind load case, Kz equals 1.13 and Gt equals 0.844.
2
Wind pressure = QK z K zt (V100 ) Gt C f (2-1a)
2
= 0.00256(1.13)(1.0)( 40) (0.844)(1.0)

= 3.91 psf

Transverse Wind Loads


Force coefficients and areas are provided in the Wind loading case.
FT 1 = 3.91(1.83)(34) = 243 lb = 0.2 kips

FT 2 = 3.91(3.11)(33) = 401 lb = 0.4 kips

FT 3 = 3.91(3.21)(62) = 778 lb = 0.8 kips

5.0.5 Construction and Maintenance (Chapter 3, Section 3.1)


Wind on Wires
Wind pressure = 3 psf
108 Electrical Transmission Line Structural Loading

Shield Wire Loads


The pulling slope is 3 horizontal to 1 vertical
 1  1800 
V = (1.5)(1757 )  + (1.5)(0.262) + 1.55 (50) = 1307 lb = 1.3 kips
 3   2 

(Alt. A controls)
V = (2)(0.262)(1800) + (2)(50) = 1043 lb = 1.0 kips

(Alt. B controls)
 0.385   5
T = (1.5)(3) 1500) + (1.5)(2)(1757 ) sin   = 446 lb = 0.4 kips
 12 (  2 

Conductor Loads
1  1800 
V = (1.5)(5971)  + (1.5)(1.075) + 1.55 (200) = 4737 lb = 4.7 kips
 3   2 

(Alt. A controls)
V = (2)(1.075)(1800) + (2)(200) = 4270 lb = 4.3 kips

(Alt. B controls)
 1.165   5
T = (1.5)(3) 1500) + (1.5)(2)(5971) sin   = 1437 lb = 1.4 kips
 12 (  2 

Wind on Structure
Transverse Wind Loads

Wind pressure = (1.5)(3) = 4.5 psf

Force coefficients and areas are provided in the Wind loading case.

FT 1 = 4.5 (1.83)(34) = 280 lb = 0.3 kips

FT 2 = 4.5 (3.11)(33) = 462 lb = 0.5 kips

FT 3 = 4.5 (3.21)(62) = 896 lb = 0.9 kips



Examples 109

5.0.6 Failure Containment (Chapter 3, Section 3.1.4 and Appendix I,


Section 1.3.1)
This loading case is based on the residual static load of a broken conduc-
tor or shield wire (0 psf wind and 0 inches of radial ice at 30 °F).

Shield Wire Loads


The RSL load factor for a broken shield wire is 1.0.

Broken Wire
 1800 
V = (0.262) + 50 = 286 lb = 0.3 kips
 2 

 5
T = (1)(1628) sin   = 71 lb = 0.1 kips
 2 

 5
L = (1)(1628) cos   = 1626 lb = 1.6 kips
 2 

Intact Wire
V = (0.262)(1800) + 50 = 522 lb = 0.5 kips

 5
T = (2)(1628) sin   = 142 lb = 0.1 kips
 2 
L = 0.0 kips

Conductor Loads

1250
Ratio of the span to insulator length = = 208
6
Note: For the purposes of this example, the span is taken to be the same
as the Ruling Span.

1250
Ratio of the span to sag = = 33
37.4
From Figure I-1, the RSL load factor is 0.7.
110 Electrical Transmission Line Structural Loading

Broken Wire
 1800 
V = (1.075) + 200 = 1168 lb = 1.2 kips
 2 

 5
T = (0.7 )(5622) sin   = 172 lb = 0.2 kips
 2 

 5
L = (0.7 )(5622) cos   = 3932 lb = 3.9 kips
 2 

Intact Wire
V = (1.075)(1800) + 200 = 2135 lb = 2.1 kips

 5
T = (2)(5622) sin   = 490 lb = 0.5 kips
 2 
L = 0.0 kips

5.1 WEIGHT SPAN CHANGE WITH BLOWOUT ON INCLINED SPANS

This example compares weight spans with and without wind for the
center tower shown in Figure 5-2. The equations are shown in Section 4.5.3
Chapter 4. Wire data are from Section 5.

Figure 5-2. Weight span for center tower with inclined spans.
Examples 111

Shield Wire
No Wind
H 1550
Cv = = = 5916 ft from Section 4.5.2
wv 0.262

 
 B 

S −1  2 
X1 = − Cv sinh  
2   S  

 Cv sinh   
  2Cv  

 
 50 
 
1250  2  = 389 ft (4-5)
X1 = − 5916 sinh−1 
2   1250  
 5916 sinh  
  2 (5916) 
 

Weight Span = 2 (1250 − 389) = 1722 ft

16.1 psf Wind


H 2917
Cv = = =11134 ft from Section 4.5.2
wv 0.262

 
 B 
 
S 2 
X1 = − Cv sinh−1  
2   S  
 Cv sinh   
  2Cv  

 
 50 
 
1250 −1  2  = 180 ft (4-5)
X1 = − 11134 sinh  
2 
 
 1250  


 11134 sinh  
  2 (11134) 

Weight Span = 2 (1250 − 180) = 2140 ft(24% increase)


112 Electrical Transmission Line Structural Loading

Conductor
No Wind
H 5305
Cv = = = 4935 ft from Section 4.5.2
wv 1.075

 
 B 
  (4-5)
S −1  2 
X1 = − Cv sinh 
2   S  

 Cv sinh   
  2Cv  

 
 50 
 
1250 −1  2  = 428 ft
X1 = − 4935 sinh 
2   1250  
 4935 sinh  
  2 ( 4935) 
 

Weight Span = 2 (1250 − 428) = 1644 ft

16.1 psf Wind


H 8530
Cv = = = 7935 ft from Section 4.5.2
wv 1.075


 
 B 
 
S −1  2 
X1 = − Cv sinh 
2   S  
 Cv sinh   
  2Cv  

 
 50 
 
1250 −1  2  = 308 ft (4-5)
X1 = − 7935 sinh 
2   1250  
 7935 sinh  
  2 (7935) 
 
Weight Span = 2 (1250 − 308) = 1884 ft(15% increase)
Examples 113

5.2 TRADITIONAL CATENARY CONSTANT

This example compares weight spans to those in Section 5.1 using the
traditional catenary constant. The traditional catenary constant is based on
the resultant unit weight (wr). The catenary constant in Section 5.1 is based
on the vertical unit weight (wv). Figure 5-2 shows the upper and lower
towers and spans.

Shield Wire
16.1 psf Wind

lb
wv = 0.262
ft

 0.385  lb
wt = 16.1 = 0.517
 12  ft

2 2 lb
wr = (0.262) + (0.517 ) = 0.580
ft

TH 2917
Cr = = = 5029 ft (4-8)
wr 0.580

 
 B 
 
S 2
X1 = − Cr sinh−1   (4-5)
2   S  
 Cr sinh   
  2Cr  

 
 50 
 
1250  2  = 424 ft
X1 = − 5029 sinh−1  
2 
 
 1250  

 5029 sinh  
  2 (5029) 

Weight Span = 2 (1250 − 424) = 1652 ft


114 Electrical Transmission Line Structural Loading

Conductor
Refer to Table 5-3.

16.1 psf Wind

lb
wv = 1.075
ft

 1.165  lb
wt = 16.1 = 1.563
 12  ft

2 2 lb
wr = (1.075) + (1.563) = 1.897
ft

TH 8530
Cr = = = 4497 ft (4-8)
wr 1.897

 
 B 
 
S −1  2  (4-5)
X1 = − Cr sinh 
2   S  
 Cr sinh   
  2Cr  


 
 50 
 
1250 −1  2 
X1 = − 4497 sinh 
2   1250  
 4497 sinh  
 2 ( 4497 )  
  

= 446 ft

Weight Span = 2 (1250 − 446) = 1608 ft


The traditional catenary constant underestimates the vertical load on


the upper tower and overestimates the vertical load on the lower tower.
APPENDIX A
DEFINITIONS, NOTATIONS, AND SI
CONVERSION FACTORS

A.1 GENERAL DEFINITIONS

Conductor Creep: Permanent elongation of conductors under everyday


tension conditions (Aluminum Company of America 1961).
MRI: Mean Recurrence Interval (years), also known as “Return Period.”
The inverse of the probability of exceedance of an environmental load
(i.e., wind, ice) in any given year. For example, a design event with the
probability of exceedance of 0.01 (1%) is associated with an MRI of 100
years.
Longitudinal: Local axis of structure that is, for tangent structures, parallel
to the direction of the conductors and perpendicular to the vertical axis.
In angle structures, this is generally the direction perpendicular to the
angle bisector.
Transverse: Local axis of structure that is, for tangent structures, perpen-
dicular to the direction of the conductors and to the vertical axis. In
angle structures, this is the direction of the angle bisector.
Vertical: Local axis of structure that is parallel with the direction of gravita-
tion force.

A.2 DEFINITIONS OF STRUCTURE TYPES

Tangent Structures: Structures having minimum line deflection angle,


typically less than 2°, such that transverse loads resulting from conduc-
tor tension are relatively small compared to those resulting from other
sources. Structures which primarily resist the vertical weight of the
wires with any accumulated ice and the transverse loads due to wind.

115
116 Electrical Transmission Line Structural Loading

Tangent structures typically utilize suspension conductor connections;


however, other means of attachment, such as post insulators, are
common.
Angle Structures: Structures marking changes in line angle along the trans-
mission line which support the same vertical and transverse loads as
tangent structures, as well as transverse loads resulting from wire ten-
sions being applied at an angle.
• May be similar to tangent structures, using suspension or post
insulators to support the conductors and transfer wind, weight, and
line angle loads to the structure.
• May be similar to strain or dead-end structures, using insulators in
series with the conductors to bring wind, weight, and line angle
loads directly into the structure.
Dead-End Structures: Structures designed as termination points for wires
and capable of supporting the loads resulting from the removal of all
wires from one or more spans.
Strain Structures: Structures similar to dead-end structures, where wire
tensions are transferred directly to the structure but are capable of resist-
ing only unbalanced/differential tensions.

A.3 DEFINITIONS OF SPAN

Span: Unless otherwise stated, span usually refers to the distance between
two adjacent structures, generally measured horizontally (Figure A-1).
Ahead Span: The span in front (generally in the direction of increasing
stationing or ascending structure numbering) of the structure in ques-
tion. In Figure A-1, Span 2 is the ahead span of Structure 11.
Back Span: The span behind (generally in the direction of decreasing sta-
tioning or descending structure numbering) the structure in question.
In Figure A-1, Span 1 is the back span of Structure 11.
Sag: The distance measured vertically from a conductor to the straight line
joining its two points of support.
Slack: The amount of conductor length difference between a straight line
made by two adjacent supports and a sagging conductor.
Weight Span: The horizontal distance between the low point of sag of
adjacent spans. It is used in calculating the vertical load the conductor
imposes on the supporting structure (Figure A-1). This may also be
referred to as the vertical span.
Wind Span: The mathematical average of the back span and the ahead
span. It is used in calculating the wind load the conductor imposes on
the supporting structure. This may also be referred to as the horizontal
span or the transverse span.
Appendix A 117

Figure A-1. Span usually refers to the distance between two adjacent structures.

A.4 NOTATION

Unless otherwise stated, the following notation is used in this manual:

A Solid tributary area of surfaces projected normal to the wind


Aml Area of all members on the longitudinal face of the structure
(Figure A-2)
Amt Area of all members on the transverse face of the structure
(Figure A-2)
Ao Area of the outline of the windward face of the structure
A s Projected area of marker ball
Bt Dimensionless response term corresponding to the quasi-static
background wind loading on the structure
Bw Dimensionless response term corresponding to the quasi-static
background wind loading on the wire (conductor or shield wire)
BWL Broken wire load
Cexp Turbulence intensity constant
Cf Force coefficient associated with the windward face of a
component
Cfl Force coefficient associated with the longitudinal faces of the
structure
118 Electrical Transmission Line Structural Loading

Figure A-2. Plan view: Structure longitudinal and transverse axis

Cft Force coefficient associated with the transverse faces of the


structure
C&M Construction and maintenance loads
COVR Coefficient of variation of component strength
d Diameter of wire (conductor or shield wire)
ds Projected diameter of wire (conductor or shield wire) with ice
accretion as appropriate
Di Straight-line distance between the supports
Dj Practical jet diameter
E Modulus of elasticity
EDT Everyday wire tension
F Wind force
f Structure/member natural frequency
Fd Force in direction of wind
Fl Wind force in the longitudinal direction
Ft Wind force in the transverse direction
ft Fundamental frequency of the free-standing structure in the
transverse direction
fw Fundamental frequency for horizontal sway of the conductor or
shield wire
FC Failure containment loads
G Gust response factor
Gt Gust response factor for the structure
Gw Gust response factor for the wire (conductor or shield wire)
H Height of hill or escarpment relative to the upwind terrain
h Insulator length
HA Horizontal line angle
Appendix A 119

HIW High intensity wind


Iz Turbulence intensity at effective height of structure
Kz Wind pressure exposure coefficient, also known as terrain factor
at height z aboveground
Kzt Topographic factor
K1 Factor to account for shape of topographic feature and maximum
speed-up effect
K2 Factor to account for reduction in speed-up with distance
upwind of downwind effect
K3 Factor to account for reduction in speed-up with height above
local terrain
L Unit length of conductor
l Length of wire from low point to support
Lh Distance upwind of crest to where the difference in ground
elevation is half the height of hill or escarpment
Lm Length of member
Ls Transverse integral scale of turbulence
LT Transverse structure load
LV Vertical structure load
LC Load case
mi Unit mass of typical ice sample
N Period of time
PDF Probability density function of a random variable
Q Air density coefficient
Qi Numerical constant to convert radial ice thickness to weight
QMRI Reliability adjustment factor
R Radial polar coordinate of downburst
r Radius of an aerial marker ball
Re Reynolds number
Rn Nominal value of component strength
RSL Residual static load for the broken wire loading condition
Rt Dimensionless resonant response term of the structure
Rw Dimensionless resonant response term of the wire
RX Wire force in the longitudinal direction
RY Wire force in the transverse direction
S Span length of the wires (conductor and ground wire)
s Member diameter or width normal to the wind
St Strouhal number
T Transverse load due to wind
t Nominal ice thickness
TH Horizontal component of tension
tz Design ice thickness for heights z above ground
V c Velocity of concurrent wind speed with ice
Vcr Critical vortex-induced wind speed
120 Electrical Transmission Line Structural Loading

Veqc Equivalent uniform velocity profile for wires


Veqt Equivalent uniform velocity profile for structures
Vi Volume of ice accreted on aerial marker ball
VJ Velocity of dropping cold air jet in a downburst model
Vmean Mean hourly wind speed
VMRI Basic wind speed for select MRI, 3-second gust at 33 ft (10 m)
height in open country
VMRI Basic wind speed, converted to hourly average
VRD Downburst outflow velocity radial component
VVR Downburst outflow velocity vertical component
V100 Basic wind speed at a 100-year mean recurrence interval,
3-second gust
VA Vertical line angle
w Wire weight per unit length
wc Unit weight of wire
WG Tornado gust width
wh Unit transverse wire load
Wi Weight of glaze ice
Wp Wind pressure
w r Resultant unit weight
wt Unit weight, wind component
wv Vertical unit weight
XL Distance from low point of sag to lower support
XR Distance from low point of sag to upper support
z Height above ground
zg Gradient height
zh Effective height above ground of the wire (conductor or ground
wire) or structure
α Power law exponent for gust wind
a Power law exponent for mean hourly wind
γ Load factor applied to weather-related loads
ε Coefficient for separation of the wire and structure response
terms in the general gust response factor equations
ζt Ratio of calculated structure damping to critical structure
damping
ζw Ratio of calculated wire damping to critical wire damping
κ Surface drag coefficient
ρ Mass density of air
ρi Ice density
Φ Solidity ratio (Am/Ao) or strength reduction factor
Ψ Angle of yaw
Appendix A 121

A.4 SI CONVERSION FACTORS

1 ft = 0.3048 meter (m)


1 in. = 25.4 millimeters (mm)
1 pound (lb) force = 4.45 Newtons (N)
1 lb/ft = 14.6 N/m
1 lb/ft2 (psf) = 47.8 Pascals (Pa) (N/m2)
1 pcf = 0.016 gram/cubic centimeter (g/cm3)
1 mile per hour (mph) = 0.45 meter/second (m/s)
APPENDIX B
RELIABILITY-BASED DESIGN

Additional information on Reliability-Based Design (RBD) methodol-


ogy, as applicable to transmission line structures, can be found in ASCE
Manual of Practice No. 111, Reliability-Based Design of Utility Pole Structures
(2006).

123
APPENDIX C
AIR DENSITY COEFFICIENT, Q

The air density coefficient, Q, converts the kinetic energy of moving air
into potential energy of pressure. The value of Q can be determined from
Equation (C-1)

Q = 0.5ρ (C-1)

where ρ is the mass density of air.


The standard value of Q based on the specific weight of air at 59  °F
(15 °C) at sea level pressure of 14.7 psi (101.325 kPa) is 0.00256 for use with
customary units (0.613 for use with SI units). For customary units, the
dimensions of Q are associated with wind speed in miles per hour and
pressure in pounds per square foot. For SI units, the dimensions of Q are
associated with wind speed in meters per second and pressure in pascals.
The use of any other value for Q for a nonstandard temperature or eleva-
tion should be based on sound engineering judgment with sufficient mete-
orological data available to justify a different value for a specific design
application.
The specific weight of air varies with temperature and atmospheric pres-
sure. Table C-1 shows values of the air density as a function of air tempera-
ture and elevation above sea level. The effect of moisture or variation in
relative humidity is assumed to be negligible.

125
126 Electrical Transmission Line Structural Loading

Table C-1.  Air Density Coefficient, Q

Air Elevation above sea level (ft)


temperature
(°F) 0 2,000 4,000 6,000 8,000 10,000
100 0.00238 0.00221 0.00205 0.00191 0.00177 0.00165
80 0.00246 0.00229 0.00213 0.00198 0.00184 0.00171
60 0.00256* 0.00237 0.00221 0.00205 0.00191 0.00178
40 0.00266 0.00247 0.00230 0.00214 0.00199 0.00185
20 0.00277 0.00257 0.00239 0.00223 0.00207 0.00192
0 0.00289 0.00268 0.00249 0.00232 0.00216 0.00201
−20 0.00293 0.00281 0.00261 0.00243 0.00226 0.00210
−40 0.00317 0.00294 0.00273 0.00254 0.00237 0.00220
Source: Adapted from Brekke (1959).
* Recommended value
APPENDIX D
CONVERSION OF WIND SPEED
AVERAGING TIME

It is recognized that wind speed values for a given record depend on the
averaging time used in the measurement of the wind speed statistics. The
use of a shorter averaging time results in a higher wind speed, whereas a
longer averaging time results in a lower wind speed. This is due to the
natural gusts and calms in wind patterns. It is often necessary to obtain
equivalent wind speeds based on different averaging periods. Conversion
of a wind speed to that representative of another averaging time can be
accomplished using the relationship shown in Figure D-1. This graph, pre-
pared from results by Durst (1960), gives the ratio, (Vt/V1-hour), of probable
maximum wind speed averaged over t seconds to hourly mean wind speed
for Exposure Category C. Note that in the graph, the value V3600 represents
the wind speed averaged over 3,600 seconds (1 hour). Additional discus-
sion of the Durst gust factor curve can be found in Miller (2011).
The hurricane simulation technique (ASCE 2017) used to develop the
wind speed maps referred to in this manual apply the relationship of gust
wind speeds described in Engineering Science Data Unit (ESDU) (1982,
1983), which have been validated for hurricane winds by Vickery and
Skerlj (2005) and Jung and Masters (2013).
The calculation of the resonant component of the dynamic response for
the gust response factor (see Appendix F) requires the mean hourly wind
speed to be calculated from the 3-second gust wind speed. The following
is an example of converting a 3-second gust wind speed (such as those
specified in Figure 2-1) to a mean hourly wind speed. This relationship is
also given in Equation (F-18) in Appendix F.
• Step 1. Obtain V3sec from Figure 2-1. This is the wind speed averaged
more than 3 seconds at a height of 33 ft (10 m) in open country terrain.

127
128 Electrical Transmission Line Structural Loading

• Step 2. Obtain the ratio (Vt/V1-hour) for t = 3 seconds from Fig-


ure D-1. This ratio is commonly taken to be 1.52.
• Step 3. Calculate the mean hourly wind speed, Vmean, based on
V3 sec
Vmean = .
1.52
For example, if Figure 2-1 indicated that a 3-second gust wind speed of
90 mph was to be used, the corresponding mean hourly wind speed is
approximately 59 mph. Note that the example provided is specific for con-
version from a 3-second gust, although the conversion to other averaging
times is carried out in a similar fashion using other values from Figure D-1.
Additional relationships (e.g., IEC 2003) have been developed to address
conversion of wind speed averaging time, and often yield varying results.
This is particularly true for Exposures B and D, where only limited histori-
cal data are available for extreme wind events. Due to the uncertainties
related to wind and its measurement, it is recommended that the conver-
sion of wind speed averaging time should only be applied to wind speeds
associated with open country terrain exposure (Exposure Category C).
If conversions of wind speed averaging times based on other exposures
are of interest to the user, the IEC 60826 (IEC 2003a, b) provides different
conversion factors based on terrain exposure categories. Alternate values
may be used for conversion of wind speed averaging time for Exposures
B or D if reasonable values can be determined from the use of local wind
data.

Figure D-1. Conversion relationship for wind speed averaging time, Exposure C.
Source: ASCE (2017).
APPENDIX E
SUPPLEMENTAL INFORMATION ON
STRUCTURE VIBRATION

E.1 INTRODUCTION

A transmission line has a large number of structures located at sites with


varying environmental and geographical exposures. Given the long length
of transmission lines, the probability of transmission structures or their
components being located in an environment prone to inducing vibration
is greater than that for non-transmission line structure types.
Vibration of a transmission structure can consist of complete structure
vibration modes, structure component vibration modes, or individual
member vibration modes. The initiation of these modes can be caused by
vibration forces induced either by the wind acting directly on the structure
or the conductor and overhead ground wires. Although general precau-
tions during the initial design of transmission structures can be considered
to reduce the possibility of vortex-induced vibration problems (ASCE
1961), the occurrence of vibration problems is highly dependent on
steady-state wind, terrain, orientation of the wind to the member cross
section, and local conditions.

E.2 STRUCTURE VIBRATION DUE TO WIND ON STRUCTURE

The majority of observed cases of member vibration can be attributed


to the vortex shedding phenomenon (e.g., von Kármán vortex street). In
this phenomenon, vortices are shed from the member in an alternating
periodic pattern, which results in oscillating forces on the member in the
plane perpendicular to the wind. If a natural frequency of the member
aligns with the frequency associated with vortex shedding, then a

129
130 Electrical Transmission Line Structural Loading

condition of harmonic resonance can occur. This is commonly referred to


as vortex-induced vibration (VIV), and may occur for a single member or
the entire structure.
Vortex shedding is related to the shape and size of the member, the
frequency of the member, wind speed, and wind orientation. The shape of
the member is taken into account through the use of the Strouhal number,
St, which is a nondimensional parameter. Based on the frequency of the
member, the corresponding Strouhal number, and the dimension of the
member exposed to the wind, a wind speed at which vortex shedding can
occur for the section can be calculated. This is referred to as the critical
wind speed and is shown in Equation (E-1):

f s
Vcr = (E-1)
St
where
Vcr = Critical wind speed associated with vortex sheading [ft/s
(m/s)],
f = Structure or member natural frequency (Hertz),
St = Strouhal number, and
s = Across-wind dimension [ft (m)].

Standard structural shapes have an average Strouhal number of 0.14.


Strouhal numbers for a variety of structural shapes can be found in Simiu
and Scanlan (1996). The structure’s natural frequency can be determined
using structural dynamic theory (Clough and Penzien 1975, Mathur et al.
1986, Paz 1980, Trainor et al. 1984). Vortex-induced motion can cause
flexural, torsional, or coupled flexural-torsional vibration modes.
Tubular steel pole arms and/or crossarms may be susceptible to damage
from vortex-induced vibration. The symmetrical shape of the arms and the
absence of the vibration damping effect of attached conductors, ground
wires, and assemblies may result in structural damage in the form of
fatigue cracking and complete failure, most commonly in relatively low
wind velocities. Cyclical stresses at the support due to vortex shedding of
the arm can exceed the endurance limits in a relatively short time period.
This cyclic loading can lead to cracking and premature failure. When
stringing operations do not occur soon after installation of tubular arms,
the installation of a weight at the end of the arm or tying the end of the arm
to the pole are common practices to reduce possible vortex-induced vibra-
tion. ASCE 48-11 (2011) contains additional information on vibration of
unloaded arms and common remedial measures to dampen oscillations.
A latticed tower, in general, presents a complex aerodynamic shape to
the wind such that consistent vortex shedding causing complete oscillation
of the structure over a prolonged period is unlikely. Therefore, only indi-
Appendix E 131

vidual member behavior of latticed towers subjected to vortex shedding


and aeroelastic instability has been studied (Modi and Slater 1983, Ward-
law 1967). Structural shapes (Thrasher 1984) and cable components, such
as guy wires and cable tension members, can be excited by vortex-induced
vibration. Latticed tower members, which are long and flexible, are par-
ticularly susceptible to vortex-induced vibration.
Large latticed structures for heavy angle sites and river crossings fre-
quently make use of stitched double-angle members. These members have
been found to be very susceptible to torsional flutter in moderate winds.
For these double angles, slenderness ratios should not exceed 200. Vibra-
tion can be reduced using wind spoilers. One example of a wind spoiler
for this application is the insertion of small, flat plates that project beyond
the horizontal legs of the angle, commonly referred to as splitter plates.
Utilities have experienced failures of the end connection plates or coped
connecting members from this wind-induced action. This phenomenon has
been documented by wind tunnel tests.
Solutions to vortex-induced problems developing during the service life
of the tower can consist of changing the frequency of the member, increas-
ing the damping, or by changing the aerodynamic properties of the mem-
ber. Changing the member cross section, or member boundary conditions
(connections), or adding intermediate structural bracing can modify the
member frequency.
Although infrequent in transmission structures, aeroelastic instability
of certain structural shapes can be a potential problem. Wind forces acting
on a structural shape that is inherently unstable at certain wind angles
cause this wind-induced vibration. Additional information can be found
in Houghton and Carruthers (1976), MacDonald (1975), Modi and Slater
(1983), Sachs (1972), Simiu and Scanlan (1996), and Slater and Modi (1971).

E.3 STRUCTURE VIBRATION DUE TO WIRE MOTION

Structure or member vibration can occur from conductor motion (Aeo-


lian, sub-conductor oscillation, and galloping) when the frequency of the
vibrating conductor corresponds to one of the natural frequencies of the
structure or its individual member(s). Approximate natural frequencies of
conductor vibration for Aeolian motion are 3 to 150 Hertz; for sub-conductor
oscillation, 0.15 to 10 Hertz; and for galloping, 0.08 to 3 Hertz (EPRI 1979).
In most instances conductor systems can be designed using dampers and
spacer dampers to prevent and/or reduce the effect of wind-induced vibra-
tion behavior. Although Aeolian vibration of wires has caused some
instances of fatigue failures of structure or hardware elements, galloping
wires have the potential to cause the most damage to the supporting struc-
ture (Brokenshire 1979, Gibbon 1984, White 1979). Latticed steel running
132 Electrical Transmission Line Structural Loading

angle suspension towers, guyed-mast dead-end structures, heavy angle


towers, and flexible “narrow-base” pole structures have been reported to
be more susceptible to damage caused by conductor galloping motion.
Field investigations have been conducted to study methods of suppressing
conductor galloping (Pohlman and Havard 1979, Richardson 1983).
APPENDIX F
EQUATIONS FOR GUST RESPONSE FACTORS

F.1 INTRODUCTION

The gust response factor (GRF) accounts for the load effects due to wind
turbulence and dynamic amplification of flexible structures and cables. It
represents the cumulative effect of the time-varying and spatially varying
fluctuating wind speeds over the range of span lengths of typical transmis-
sion lines, as well as the effect of the wind on supporting structures. The
approach for the gust response factors provided in Section 2.1.5 of Chap-
ter 2 of this manual are based on work by Davenport (1979) for estimating
the peak response of transmission line systems to gusting winds, as well
as wind loading provisions in ASCE 7-16 (2017).
The original Davenport GRF equations were developed using statistical
methods which involve the spatial correlation and energy spectrum of tur-
bulent wind, as well as the dynamic characteristics of transmission line
components. The complete GRF equations include amplification factors
that account for the resonant component of the dynamic response of struc-
tures and wires. The derivation of the gust response factor is given in Dav-
enport (1979), and their application to typical towers and wires is discussed
in previous versions of this manual (ASCE 1991, 2010a). In the second
edition of this manual (ASCE 1991), simplified equations are presented
where the resonant component of the dynamic response is negligible for a
large range of typical tower configurations. These simplifications were
based on a theoretical appraisal of transmission line behavior, as well as an
assessment of available full-scale data. The underlying assumptions and
limitations of the simplified procedure are discussed further in Section F.4.
In the third edition of this manual (ASCE 2010a), the gust response factors

133
134 Electrical Transmission Line Structural Loading

were modified to be made compatible with the 3-second gust wind speed
for consistency with ASCE 7.
The equations for some of the components of the gust response factor
have been modified for this edition of the manual. Most notable are the
removal of the parameters κ (surface drag coefficient) and E (exposure fac-
tor); these have been replaced with more current parameters used in the
description of atmospheric boundary layer wind. As well, the exponents
for the power law now reflect a mean hourly wind speed for which the
equations were derived. This is in contrast to the use of power law expo-
nents for the fastest-mile wind speed used previously. The updated
approach involves the calculation of the turbulence intensity of the wind
at the effective height of the structure or wires, as well as the use of separate
peak factors for the background and resonant components of the dynamic
response. The revisions to the methodology reflect the state of the art in the
calculation of wind loads on structures and are consistent with the meth-
odology applied in ASCE 7. A detailed description of these changes is given
by Mara (2015).
These equations are based on idealized conditions that may or may not
reflect the true weather events that a transmission line structure may expe-
rience. Thus, the results obtained by the application of these equations
within this context should be considered approximate. The purpose of this
appendix is to present the gust response factor equations and define the
various wind, exposure, and dynamic parameters used. The approach for
the gust response factor of structures and lines is consistent with that
developed by Davenport (1979); however, some of the nomenclature has
been slightly modified to incorporate the relationships used in the develop-
ment of the ASCE 7-16 wind load criteria that form the basis of this
manual.
The equations are given in this appendix without derivation. However,
interested readers may refer to several papers that have dealt with this
subject (e.g., Davenport 1962, 1967, 1977, 1979; Vellozzi and Cohen 1968).

F.2 NOMENCLATURE

The following nomenclature is used in this appendix:

Bt Dimensionless response term corresponding to the quasi-static


background component of the dynamic response of the tower
[Equation (F-6)]
Bw Dimensionless response term corresponding to the quasi-static
background component of the dynamic response of the wires
[Equation (F-14)]
Appendix F 135

cexp Turbulence intensity constant, based on exposure (Table F-1)


Cf Force coefficient for the wires (typically taken as 1.0; see Appen-
dix G)
ds Diameter of the wire (conductor or shield wire) in inches
ft Fundamental frequency of the tower or structure in the trans-
verse direction, in Hz (see Table F-2 for approximate values)
fw Fundamental frequency for horizontal sway of the conductor or
shield wire, in Hz [Equation (F-20)]
gB Peak factor for the background component of the dynamic
response (same for tower and wires, constant value of 3.6)
gRt Peak factor for the resonant component of the dynamic response
of the tower [Equation (F-4)]
gRw Peak factor for the resonant component of the dynamic response
of the wires [Equation (F-12)]
gv Peak factor for the turbulence of the wind (constant value of 3.6)
Gt Gust response factor for wind loading on structure
[Equation (F-1)]
Gw Gust response factor for wind loading on wires [Equation (F-9)]
Iz Turbulence intensity at effective height of the tower/structure or
wire [Equations (F-2) and (F-10)]
Ls Integral length scale of turbulence (ft) (Table F-1)
Rt Dimensionless resonant response term of the structure
[Equation (F-7)]
Rw Dimensionless resonant response term of the wires
[Equation (F-15)]
S Design wind span (ft)
Sag Wire sag at midspan (ft)
VMRI Basic wind speed, 3-second gust at 10 m height in open country
terrain (Figure 2-1)
VMRI Basic wind speed, converted to mean hourly wind speed
(Appendix D)
Vo Mean hourly wind speed (ft/s) at effective height of the tower/
structure or wires, based on exposure
zg Gradient height (ft) (Table F-1)
zh Effective height of wires and/or structure (Section 2.1.4.3)
a Power law exponent for mean hourly wind (Table F-1)
a Power law exponent for gust wind (Table F-1)
e Separation coefficient reflecting the non-coincidence of tower
and wire loads (for typical transmission line systems, ε is
approximated by 0.75)
ζ t Damping ratio of structure relative to critical (see Table F-2 for
approximate values)
ζw Damping ratio of wires relative to critical [Equation (F-21)]
136 Electrical Transmission Line Structural Loading

F.3 RELEVANT EQUATIONS

The gust response factors for the tower, Gt, and the wires, Gw, are given
in Equations (F-1) and (F-9), respectively. All parameters are described in
Section F.2, and wind parameters by exposure category are listed in
Table F-1. Note that the subsequent equations contain both the background
and resonant components of the dynamic response, as opposed to the sim-
plified versions provided in Section 2.1.5.1. For unique structures (e.g.,
complex structural configurations, very tall structures, structures with low
fundamental frequencies) and for very long spans, the inclusion of the
resonant component may be of importance. As the magnitude of the reso-
nant effect decreases, Equations (F-1) and (F-9) converge to Equations (F-8)
and (F-16), respectively.

Table F-1.  Wind Parameters by Exposure Category

Exposure a a cexp Ls (ft) zg (ft)

B 7.0 4 0.3 170 1200

C 9.5 6.5 0.2 220 900

D 11.5 9 0.15 250 700


Source: Adapted from ASCE 7-16 (ASCE 2017).

The gust response factor for the tower, Gt, with an effective height zh is
calculated as
 1 + 1.7 I e g 2 B2 + g 2 R 2 
 Rt t 
Gt = 
z B t
 (F-1)
 1 + 1. 7 I g 
 z v 
1
 33 6
in which I z = cexp   (F-2)
 z 
h

gb = gv = 3.6 (F-3)

0.577
g Rt = 2ln(3600 ft ) + (F-4)
2ln(3600 ft )

e = 0.75 (F-5)
Appendix F 137

1
Bt = (F-6)
0.56 zh
1+
Ls

−5
 0.0123  ft zh  3
Rt =    (F-7)
 z  V 
t o

Note that if the resonant component is neglected, Equation (F-1) con-


verges to the simplified expression presented in Chapter 2:

 1 + 4.6 I z Bt 
Gt =   (F-8)
 1 + 6.1I z 

The gust response factor for the wires, Gw, with an effective height zh is
calculated as

 1 + 1.7 I e g 2 B2 + g 2 R2 
  (F-9)
Gw = 
z B w Rw w

 1 + 1. 7 I g 
 z v

1
 33 6
in which I z = cexp   (F-10)
 z 
h

gb = gv = 3.6 (F-11)

0.577
g Rw = 2ln(3600 f w ) + (F-12)
2ln(3600 f w )

e = 0.75 (F-13)

1
Bw = (F-14)
0.8S
1+
Ls
138 Electrical Transmission Line Structural Loading

−5
 0.0113  zh  f w zh  3
Rw =     (F-15)
 zw
 S  V 
o

Note that if the resonant component is neglected, Equation (F-9) con-


verges to the simplified expression presented in Chapter 2:

 1 + 4.6 I z Bw 
Gw =   (F-16)
 1 + 6.1I z 

For the calculation of Rt and Rw, the resonant components of the dynamic
response, the mean hourly wind speed (ft/s) at the effective tower and wire
heights is calculated as

1
 z a  88 

Vo = 1.66  h   VMRI (F-17)
 z g   60 

VMRI
where VMRI = (F-18)
1.52
Approximate ranges in the fundamental natural frequency and damp-
ing ratio for suspension structures are given in Table F-2. The frequencies
in this table are based on a limited review of typical suspension structure
dynamic properties and are not intended to be applicable for every type of
transmission structure. Since little data are available on damping ratios for
transmission line structures, the values given in Table F-2 are conservative
estimates for most structure types. The designer is encouraged to perform
numerical analyses or dynamic tests in order to determine the appropriate
properties. Additional information on the dynamic response of latticed
towers and guyed masts can be found in ASCE (2002).

Table F-2.  Approximate Dynamic Properties for Transmission Structures

Fundamental frequency
Type of structure (Hz), ft Damping ratio, zt

Latticed tower 2.0–4.0 0.04


H-frame 1.0–2.0 0.02
Pole 0.5–1.0 0.02
Appendix F 139

Based on a past survey of transmission latticed tower frequencies, the


frequency of a tower can be estimated as
328
ft = (F-19)
h
where h is the total height of the tower (ft).

The fundamental frequency of pole-type structures can be calculated


using general engineering theory. The taper dimension must be included
in the estimation of the fundamental frequency.
The following equations may be used to approximate the frequency of
the wires and the damping ratio of the wires if they are not known:

1
fw = (F-20)
sag

 12V 
z w = 0.000048C f  o
 (F-21)
 f w d 

It should be noted that Equations (F-19), (F-20), and (F-21), as well as the
values in Table F-2, should be regarded as estimates only. If more accurate
estimates of frequency or damping are available, such as those obtained
through numerical analyses or dynamic tests, these values may be used to
improve the estimation of the gust response factors.

F.4 ASSUMPTIONS AND LIMITATIONS

To derive the equations in Chapter 2, Section 2.1.5 (Eq. 2-4 and Eq. 2-5),
some simplifying assumptions were made based on work carried out by
Davenport (1979). These assumptions are listed below:
1. The separation coefficient, e, is equal to 0.75 and reflects the nonco-
incident nature of strong wind loads on the structures and wires.
2. The statistical peak factors for background response and wind
loading, gB and gv, are equal to 3.6. The peak factors are approxi-
mated for structures responding to buffeting wind with a broad
spectrum of energy over a range of frequencies.
3. The resonant component of the dynamic response for both structure
and wire systems, Rt and Rw, can be neglected for transmission
structures of typical size. For typical systems, tower vibration is
small due to the relatively high frequency of the structure. Wire
140 Electrical Transmission Line Structural Loading

vibration is generally low due to the high aerodynamic damping


associated with wire motion at design wind speeds. These aspects
are reflected in the design equations and supported by the observa-
tion that the resonant component is often not significant in trans-
mission lines. For tall or unique structures, designers are
encouraged to consider including the resonant component of the
response in the design.

F.5 EXAMPLE

The following example calculates the gust response factors for the exam-
ple structure and line given in Chapter 5 of this manual. The gust response
factors for the structure and the wires are calculated for (1) the background
dynamic response only (simplified method in Chapter 2), and (2) the com-
plete dynamic response (background and resonant).
The effective height of the tower is 59.3 ft, and the effective height of the
wires (this example considers the conductors only) is 74 ft (conservative
approach based on strong wind conditions). The total height of the tower
is 89 ft, and the wind span is 1,500 ft. The diameter of the conductors is
1.165 inches, and the estimated sag of the conductor is 36 ft. A structural
damping of 0.03 (3%) is assumed for the tower. The basic wind speed for
the example in this appendix is V100 = 96 mph. The tower is in terrain char-
acteristic of Exposure Category C.

Structure Gust Response Factor, Gt

1. Background dynamic response only.


The equation for the simplified gust response factor [Equation (F-8)] is

 1 + 4.6 I z Bt 
Gt =  
 1 + 6.1I z 

From Table F-1, the wind parameters for Exposure C are

Table F-1.  Wind Parameters by Exposure Category.

Exposure a a cexp Ls (ft) zg (ft)

C 9.5 6.5 0.2 220 900


Appendix F 141

1
 33 6
From Equation (F-2) I z = cexp  
 z 
h

1
 33 6
= 0.2 
 59.3 

= 0.18

From Equation (F-3) gb = gv = 3.6

From Equation (F-5) e = 0.75

1
From Equation (F-6) Bt =
0.56 zh
1+
Ls

1
=
0.56(59.3)
1+
220
= 0.932

 1 + 4.6 I z Bt 
From Equation (F-8) Gt =  
 1 + 6.1I z 

 1 + 4.6(0.181)(0.932) 
=  
 1 + 6.1(0.181) 

= 0.844

2. Consider the complete dynamic response.


The equation for the detailed gust response factor [Equation (F-1)] is

 1 + 1.7 I e g 2 B2 + g 2 R 2 
 
Gt =
z B t Rt t
 
 1 + 1. 7 I g
z v 
142 Electrical Transmission Line Structural Loading

The parameters Iz, e, gB, gv, and Bt are as for the background response.
From Equation (F-19), the frequency of the tower can be estimated based
on a full tower height, h , of 89 ft
328 328
ft = = = 3.69 Hz
h 89

zt = 0.03

0.577
From Equation (F-4) g Rt = 2ln(3600 ft ) +
2ln(3600 ft )

0.577
= 2ln(3600 × 3.69) +
2ln(3600 × 3.69)

= 4.49

VMRI 96
From Equation (F-18) VMRI = = = 63.2 mph
1.52 1.52

1
 z a  88 

From Equation (F-17) Vo = 1.66  h   VMRI
 z g   60 

1
 59.3 6.5  88 
= 1.66  
 900   60 (
63.2)

= 101.3 ft/s
−5
 0.0123  ft zh  3
From Equation (F-7) Rt =   
 z  V 
t o

−5
 0.0123  (3.69)(59.3)  3
=  
 0.03  101.3 

= 0.337
Appendix F 143

Substituting solved values in Equation (F-1)

 1 + 1.7 I e g 2 B2 + g 2 R 2 
 
Gt = 
z B t Rt t

 
 1 + 1.7 I z g v 

æ 2 2 ö÷
çç 1 + 1.7(0.181)(0.75) (3.6) (0.932) + (4.49)2 (0.337)2 ÷÷
= çç ÷÷
ç
çè 1 + 1.7 (0.181)(3.6) ÷÷
ø
= 0.877

Note that, in this example, the gust response factor considering the com-
plete dynamic response is about 4% greater than that calculated with the
simplified equations in Chapter 2 (which consider the background compo-
nent only).

Wire Gust Response Factor, Gw

1. Background dynamic response only


The equation for the simplified gust response factor [Equation (F-16)] is

1
 33 6
From Equation (F-10) I z = cexp  
 z 
h

1
 33 6
= 0.2  
 74 

= 0.175

From Equation (F-11) gb = gv = 3.6

From Equation (F-13) e = 0.75

1
From Equation (F-14) Bw =
0.8S
1+
Ls
144 Electrical Transmission Line Structural Loading

1
=
0.8(1500)
1+
220
=0.394

 1 + 4.6 I z Bw 
From Equation (F-16) Gw =  
 1 + 6.1I z 

 1 + 4.6(0.177 )(0.394) 
=  
 1 + 6.1(0.177 ) 

= 0.637

2. Consider the complete dynamic response.


The equation for the detailed gust response factor [Equation (F-9)] is

 1 + 1.7 I e g 2 B2 + g 2 R2 
 
Gw = 
z B w Rw w
 
 1 + 1. 7 I g
z v 

The parameters Iz, e, gB, gv, and Bw are as for the background response
only.
The frequency of the conductor wire can be estimated based on a sag of
the wire of 36 ft

1 1
fw = = = 0.167
sag 36

From Equation (F-21), the damping of the conductor wire can be esti-
mated based on wire and wind parameters

1
 z a  88 

Vo = 1.66  h   VMRI
 z g   60 
1

 74 6.5  88 
= 1.66   63.2)
 900   60 (

= 104.8 ft/s
Appendix F 145

 12V 
z w = 0.000048C f  o

 f w d 

 12(104.8) 
= 0.000048(1.0) 
 (0.167 )(1.165) 

= 0.310

0.577
From Equation (F-12) g Rw = 2ln(3600 f w ) +
2ln(3600 f w )

0.577
= 2ln(3600 × 0.167 ) +
2ln(3600 × 0.167 )

= 3.74

−5
 0.0113  zh  f w zh  3
From Equation (F-15) Rw =    
 z  S  V 
w o

−5
 0.0113  74  (0.167 )(74)  3
=   
 0.310  1500  104.8 

= 0.252

Substituting solved values in Equation (F-9)

 1 + 1.7 I e g 2 B2 + g 2 R2 
 
Gw = 
z B w Rw w
 
 1 + 1. 7 I g
z v 

 1 + 1.7(0.175)(0.75) (3.6)2 (0.394)2 + (3.74)2 (0.252)2 


 
=  
 1 + 1.7(0.1175)(3.6) 

= 0.666

Note that in this case, the gust response factor considering the complete
dynamic response is about 5% greater than that calculated with the simplified
equations in Chapter 2 (which consider the background component only).
APPENDIX G
SUPPLEMENTAL INFORMATION ON FORCE
COEFFICIENTS

G.1 CONDUCTOR AND SHIELD WIRE FORCE COEFFICIENTS

Wind tunnel test data, such as those shown in Figure G-1, indicate that
measured force coefficients for stranded wires show a wide range of varia-
tion depending on Reynolds number and the type of stranding. For this
reason, there is also a wide variation in values recommended by various
design codes and guides as illustrated in Figure G-2.
A force coefficient of 1.0 is recommended in Chapter 2, Section 2.1.6.2
for all conductors and shield wires. This is the same value recommended
in NESC (2012). The data in Figure G-1 indicate that the force coefficient
can be significantly greater than 1.0, particularly for Reynolds numbers less
than 3 × 104 (small wires under nominal wind speed). For Reynolds num-
bers above this value, the force coefficients are reduced to a value of 1.0 or
less. The expression for Reynolds number is given in Equation (2-9).
For a 0.5 inch diameter wire or larger, the Reynolds number will exceed
3 × 104 for the range of design wind speeds given in Chapter 2, Figure 2-1.
For this reason, a value of 1.0 has been chosen for all conductors and shield
wires. However, force coefficients larger than 1.0 are often appropriate,
especially on wires having a small diameter (< 0.5 inch) and wires having
accreted ice.

147
148 Electrical Transmission Line Structural Loading

Force Coefficient

Reynolds Number

Figure G-1. Force coefficients for conductors based on wind tunnel tests.
Source: Data from ASCE (1961), Birjulin et al. (1960), Castanheta (1970),
Engleman and Marihugh (1970), Richards (1965), and Watson (1955).
1.4

1.3

1.2

1.1

1.0

0.9

0.8

0.7
Force Coefficient

0.6

0.5

0.4

0.3

0.2

0.1

0.0
3 4 5 6 7 891 2 3 4 5 6 7 891 2 3 4 5 6 7 891 2 3
x 10 x 10 x 10

Reynolds Number

Figure G-2. Force coefficients for conductors based on code values.


Appendix G 149

G.2 MEMBER FORCE COEFFICIENTS

Table 2-5 lists recommended force coefficients for some common struc-
tural shapes used in transmission structures. Table G-1 lists force coeffi-
cients from various sources for these members and for additional shapes
not listed in Table 2-5. For some shapes, values are given for variations in
surface roughness, Reynolds number, corner radius ratio, yaw angle, or
test conditions.
The force coefficients of asymmetrical shapes are dependent on the ori-
entation of the wind with respect to the cross section of the member. No
general equation exists for this condition; however, values have been deter-
mined through wind tunnel testing. These instances are indicated in
Table G-1.
150 Electrical Transmission Line Structural Loading

Table G-1.  Member Force Coefficients.

PROJ. AREA = s x LENGTH

W IN D
s

Circle

Surface Reynolds number Force coefficient Reference


Any < 3.5 × 105 1.2 Scruton and
Newberry (1963)
Any < 4.1 × 105 1.2 MacDonald (1975)
Smooth — 0.7 ASCE (1990a)
5
Smooth < 10 1.0 Sachs (1978)
Smooth < 3.0 × 105 1.1 AASHTO (1975)
5
Smooth > 3.5 × 10 0.7 Scruton and
Newberry (1963)
Smooth > 4.1 × 105 0.6 MacDonald (1975)
5 5 6
Smooth 3 × 10 < Re < 6 × 10 14.5 × 10 /Re1.3 AASHTO (1975)
Smooth > 6.0 × 105 0.45 AASHTO (1975)
5
Rough > 4.1 × 10 1.2 MacDonald (1975)
Rough — 0.9 ASCE (1990a)
5
Very rough > 3.5 × 10 1.0 Scruton and
Newberry (1963)
Very rough — 1.2 ASCE (1990a)
Appendix G 151

PROJ. AREA = s x LENGTH


r = RADIUS OF CORNERS
WIND
R = RADIUS OF INSCRIBED CIRCLE
s

16-sided polygon

Corner radius (r/R) Reynolds number Force coefficient Reference


< 0.26 > 6.0 × 105 0.83–1.08(r/R) James (1976)
> 0.26 > 6.0 × 105 0.55 James (1976)

PROJ. AREA = s x LENGTH


r = RADIUS OF CORNERS
WIND
R = RADIUS OF INSCRIBED CIRCLE
s

12-sided polygon

Reynolds Force
Corner radius (r/R) number coefficient Reference
0 < 3.5 × 105 1.3 Scruton and
Newberry (1963)
0 < 8.2 × 105 1.3 MacDonald (1975)
0 > 3.5 × 105 1.0 Scruton and
Newberry (1963)
0 > 8.2 × 105 1.1 MacDonald (1975)
0.09 < r/R < 0.34 > 106 0.936–1.087(r/R) James (1976)
> 0.125 < 3.0 × 105 1.2 AASHTO (1975)
> 0.125 3.0 × 105 < Re 2,322/Re0.6 AASHTO (1975)
< 6.0 × 105
> 0.125 > 6.0 × 105 0.79 AASHTO (1975)
> 0.34 > 106 0.57 James (1976)
152 Electrical Transmission Line Structural Loading

PROJ. AREA = s x LENGTH


r = RADIUS OF CORNERS
W I ND
R = RADIUS OF INSCRIBED CIRCLE
s

8-sided polygon

Corner radius Reynolds


(r/R) number Force coefficient Reference
0 — 1.2 AASHTO (1975)
0 — 1.4 ASCE (1990a),
MacDonald
(1975)
0.09 < r/R < 0.59 > 106 1.422–1.368(r/R) James (1976)
> 0.59 > 106 0.744–0.194(r/R) James (1976)

PROJ. AREA = s x LENGTH

WIND
s

2s

Ellipse, wind on narrow side

Reynolds
Sides number Force coefficient Reference
5
Smooth < 6.9 × 10 0.7 MacDonald
(1975)
Smooth > 6.9 × 105 0.2 MacDonald
(1975)
Multi-sided — (C/3)(4 – D/d) AASHTO (1975)
where D = Major diameter
d = Minor diameter
D/d = 2.0
C = Force coefficient of cylindrical shape with
diameter equal to D
Appendix G 153

PROJ. AREA = s x LENGTH

WIND
s

s/2

Ellipse, wind on broad side

Reynolds
Sides number Force coefficient Reference

Smooth < 5.5 × 105 1.7 MacDonald


(1975)
Smooth > 5.5 × 105 1.5 MacDonald
(1975)
Multisided — 1.7(D/d – 1) + AASHTO (1975)
C(2 – D/d)

Cs

Cn
PROJ. AREA = s x LENGTH
Cn = COEFFICIENT NORMAL TO THE SURFACE
Cs = COEFFICIENT 90° TO Cn
REF
s
WIND

0.1 s

Flat plate

Angle Cn Cs Reference

0° 2.0 0.0 Scruton and Newberry


(1963), Sachs (1978)
45° 1.8 0.1 Sachs (1978)
90° 0.0 0.1 Sachs (1978)
154 Electrical Transmission Line Structural Loading

Cs

Cn

REF PROJ. AREA = s x LENGTH


s Cn = COEFFICIENT NORMAL TO THE SURFACE
WIND Cs = COEFFICIENT 90° TO Cn

r = RADIUS OF CORNERS
R = RADIUS OF INSCRIBED CIRCLE
s/2

Rectangle

Corner
radius (r/R) Angle Cn Cs Reference
0 0° 2.2 0.0 Scruton and
Newberry (1963)
0 0° 2.1 0.0 Sachs (1978)
0 45° 1.4 0.7 Sachs (1978)
0 90° 0.0 0.75 Sachs (1978)
0.08 0° 1.9 0.0 MacDonald (1975)
0.25 0° 1.6 0.0 Scruton and
Newberry (1963)

WIND
PROJ. AREA = s x LENGTH
s
r = RADIUS OF CORNERS
R = RADIUS OF INSCRIBED CIRCLE

2s

Rectangle

Corner radius Reynolds


(r/R) number Force coefficient Reference

0.0 — 1.4 Scruton and


Newberry (1963)
0.167 — 0.7 MacDonald (1975)
0.5 — 0.4 Sachs (1978)
Appendix G 155

WIND
PROJ. AREA = 1.414 x s x LENGTH
r = RADIUS OF CORNERS
R = RADIUS OF INSCRIBED CIRCLE

Square, wind at apex (cornering)

Corner radius Reynolds


(r/R) number Force coefficient Reference

0.0 — 1.5 ASCE (1990b),


Scruton and
Newberry (1963)
0.33 < 6.86 × 105 1.5 MacDonald (1975)
5
0.33 > 6.86 × 10 0.6 MacDonald (1975)

WIND
PROJ. AREA = s x LENGTH
s
r = RADIUS OF CORNERS
R = RADIUS OF INSCRIBED CIRCLE

Square, wind at side

Corner radius Reynolds


(r/R) number Force coefficient Reference

0.0 — 2.0 ASCE (1990b),


Scruton and
Newberry (1963)

0.167 < 6.86 × 105 1.3 MacDonald (1975)


0.167 > 6.86 × 105 0.6 MacDonald (1975)
0.33 < 2.7 × 105 1.0 MacDonald (1975)
0.33 > 2.7 × 105 0.5 MacDonald (1975)
156 Electrical Transmission Line Structural Loading

Cs

s/2 Cn
PROJ. AREA = s x LENGTH
Cn = COEFFICIENT NORMAL TO THE SURFACE
REF Cs = COEFFICIENT 90° TO Cn
WIND s

Unequal leg angle

Angle Cn Cs Reference

0° 1.9 0.95 Sachs (1978)


45° 1.8 0.8 Sachs (1978)
90° 2.0 1.7 Sachs (1978)
135° –1.8 –0.1 Sachs (1978)
180° –2.0 0.1 Sachs (1978)

Cs

0.48 s
Cn
PROJ. AREA = s x LENGTH
Cn = COEFFICIENT NORMAL TO THE SURFACE
REF Cs = COEFFICIENT 90° TO Cn
s

WIND

I-beam

Angle Cn Cs Reference

0° 2.05 0.0 Sachs (1978)


45° 1.95 0.6 Sachs (1978)
90° 0.5 0.9 Sachs (1978)
Appendix G 157

Cs

0.43 s
Cn
PROJ. AREA = s x LENGTH
Cn = COEFFICIENT NORMAL TO THE SURFACE
REF Cs = COEFFICIENT 90° TO Cn
s

WIND

Channel

Angle Cn Cs Reference

0° 2.05 0.0 Sachs (1978)


45° 1.85 0.6 Sachs (1978)
90° 0.0 0.6 Sachs (1978)
135° –1.6 0.4 Sachs (1978)
180° –1.8 0.0 Sachs (1978)

Cs

s
Cn
PROJ. AREA = s x LENGTH
Cn = COEFFICIENT NORMAL TO THE SURFACE
REF Cs = COEFFICIENT 90° TO Cn
s

WIND

Wide flange

Angle Cn Cs Reference

0° 1.6 0.0 Sachs (1978)


45° 1.5 1.5 Sachs (1978)
90° 0.0 1.9 Sachs (1978)
158 Electrical Transmission Line Structural Loading
Cs

1.6s

Cn

REF
s
PROJ. AREA = s x LENGTH
WIND Cn = COEFFICIENT NORMAL TO THE
SURFACE
Cs = COEFFICIENT 90° TO Cn

Built-up section

Angle Cn Cs Reference

0° 1.4 0.0 Sachs (1978)


45° 1.2 1.6 Sachs (1978)
90° 0.0 2.2 Sachs (1978)

Cs

s
PROJ. AREA = s x LENGTH
Cn = COEFFICIENT NORMAL TO THE SURFACE
REF Cs = COEFFICIENT 90° TO Cn
Cn

WIND

Equal leg angle

Angle Cn Cs Reference

0° 1.8 1.8 Sachs (1978)


45° 2.1 1.8 Sachs (1978)
90° –1.9 –1.0 Sachs (1978)
135° –2.0 0.3 Sachs (1978)
180° –1.4 –1.4 Sachs (1978)
Appendix G 159
Cs
0.45 s

Cn
REF
PROJ. AREA = s x LENGTH
s
Cn = COEFFICIENT NORMAL TO THE SURFACE
Cs = COEFFICIENT 90° TO Cn
WIND

Double angle

Angle Cn Cs Reference

0° 1.6 0.0 Sachs (1978)


45° 1.5 –0.1 Sachs (1978)
90° –0.95 0.7 Sachs (1978)
135° –0.5 1.05 Sachs (1978)
180° –1.5 0.0 Sachs (1978)

s Cs

Cn
PROJ. AREA = s x LENGTH
s
REF Cn = COEFFICIENT NORMAL TO THE SURFACE
Cs = COEFFICIENT 90° TO Cn

WIND

Built-up angles

Angle Cn Cs Reference

0° 1.75 0.1 Sachs (1978)


45° 0.85 0.85 Sachs (1978)
90° –0.1 1.75 Sachs (1978)
135° –0.75 0.75 Sachs (1978)
180° –1.75 –0.1 Sachs (1978)
160 Electrical Transmission Line Structural Loading

1.1 s Cs

Cn
PROJ. AREA = s x LENGTH
REF s
Cn = COEFFICIENT NORMAL TO THE SURFACE
Cs = COEFFICIENT 90° TO Cn

WIND

T-section

Angle Cn Cs Reference
0° 2.0 0.0 Sachs (1978)
45° 1.2 0.9 Sachs (1978)
90° –1.6 2.15 Sachs (1978)
135° –1.1 2.4 Sachs (1978)
180° –1.7 2.1 Sachs (1978)

G.3 ASPECT RATIO

The force coefficients given in Chapter 2, Section 2.1.6.2 and in Sec-


tion  G.2 are for infinitely long members and are applicable to members
with aspect ratios greater than 40. Adjustment factors for members with
aspect ratios less than 40 may be applied as follows (MacDonald 1975):

Cf' = (c)(Cf) (G-1)


where
c = Correction factor for aspect ratio (Table G-2),
Cf = Force coefficient from Section 2.1.6.2 or G.2, and
Cf' = Force coefficient corrected for aspect ratio.
Table G-2.  Aspect Ratio Correction Factors

Aspect ratio Correction factor (c)


0–4 0.6
4–8 0.7
8–40 0.8
> 40 1
Note: Aspect ratio = (Lm/ds) except for members attached to the ground where
aspect ratio = (2Lm/ds), in which Lm = member length and ds = member diameter
or width.
Appendix G 161

G.4 LATTICED TRUSS STRUCTURE FORCE COEFFICIENTS

The force coefficients calculated using Table 2-4 and Equation (2-13) in
Chapter 2 represent the recommended values for square-section and
triangular-section latticed structures having flat-sided and rounded mem-
bers. The recommended force coefficients, which are taken directly from
ASCE 7-16 (2017), account for the wind forces acting on the windward and
leeward faces of the latticed tower. Therefore, they are influenced by the
solidity ratio, which is defined in Equation (2-10). As the solidity ratio
increases, the force coefficient is reduced due to the shielding effect of the
members in the windward face(s) of the tower.
Figures G-3 through G-6 provide information from various other codes,
standards, and tests for force coefficients for latticed towers with wind
normal to a face. These figures are for towers having either square or tri-
angular cross-sections and comprised of flat-sided or rounded members.
Figures G-7 through G-10 provide information from various codes and
standards for force coefficients for latticed towers with yawed wind. These
figures are for latticed tower structures having either square or triangular
cross-sections and comprised of flat-sided or rounded members. Whit-
bread (1979) has published other data relating to wind forces on latticed
towers having a wide variety of shapes, solidity ratios, and wind direc-
tions. The variation of the force coefficient with yaw angle was examined
for square sections by Bayar (1986) and for a typical cross-arm section by
Mara et al. (2010).

Figure G-3. Force coefficients for square-section towers having flat-sided


members with wind normal to a face.
162 Electrical Transmission Line Structural Loading

Figure G-4. Force coefficients for square-section towers having rounded mem-
bers with wind normal to a face.

Figure G-5. Force coefficients for equilateral triangular-section towers having


flat-sided members with wind normal to a face.
Appendix G 163

Figure G-6. Force coefficients for equilateral triangular-section towers having


rounded members with wind normal to a face.

Figure G-7. Force coefficients for square-section towers having flat-sided


members with diagonal wind.
164 Electrical Transmission Line Structural Loading

Figure G-8. Force coefficients for square-section towers having rounded mem-
bers with diagonal wind.

Figure G-9. Force coefficients for equilateral triangular-section towers having


flat-sided members with cornering wind.
Appendix G 165

Figure G-10. Force coefficients for equilateral triangular-section towers having


rounded members with cornering wind.

G.5 FORCE COEFFICIENTS OF ICED COMPONENTS

When calculating forces due to wind on ice-covered wires, the force


coefficient is dependent on the shape of the ice buildup (McComber et al.
1982). However, typical force coefficients of ice-covered wires are not
known. Some organizations recommend using force coefficients other than
1.0 for wires covered with ice (IEC 2003, ISO Standard 12494).
APPENDIX H
SUPPLEMENTAL INFORMATION
ON ICE LOADING

H.1 THEORY AND CONDITIONS OF ICE FORMATION

In a general sense, the meteorological parameters that influence the type


and amount of ice that forms under different conditions are well known.
Liquid water content of supercooled clouds and precipitation intensity for
freezing precipitation icing and sticky snow determine the amount of water
available for ice formation. The ice properties are determined by the air
temperature, wind speed, drop size, and supercooled liquid water content
of clouds, fog, or precipitation intensity and type. The icing phenomenon
is best classified by the causal meteorological conditions. In the following
paragraphs, the various icing mechanisms are described as it is important
for the engineer to understand the conditions which may result in severe
loads on transmission lines.

H.1.1 Precipitation Icing


Freezing rain (or drizzle) is a common icing mechanism. Freezing rain
occurs when warm, moist air is forced over a layer of subfreezing air at the
Earth’s surface. The precipitation usually begins as snow, which melts as
it descends through the layer of warm air aloft. The drops cool as they fall
through the cold surface air layer and freeze on contact with structures, or
the ground, to form glaze ice. Upper air data indicate that the cold surface
air layer is typically between 1,000 ft (300 m) and 3,900 ft (1,200 m) thick
(Young 1978), and averages approximately 1,600 ft (500 m) (Bocchieri 1980).
The warm air layer aloft averages 5,000 ft (1,500 m) thick in freezing rain,
but in freezing drizzle the entire temperature profile may be below 32 °F
(0 °C) (Bocchieri 1980). Precipitation associated with slowly moving frontal

167
168 Electrical Transmission Line Structural Loading

systems can alternate between snow and freezing rain to form a composite
slow-glaze accretion on structures. The density of glaze is usually assumed
to be 56 to 57 pcf (900 to 917 kg/m3).
In freezing rain, the water impingement rate is often greater than the
freezing rate. The excess water starts to drip off and may freeze as icicles,
resulting in a variety of accretion shapes that range from a smooth, cylin-
drical sheath through a crescent on the windward side with icicles hanging
on the bottom to large, irregular protuberances. The shape of a glaze accre-
tion depends on the varying meteorological factors and the cross-sectional
shape of the structural member or component, its spatial orientation, and
flexibility.

H.1.2 In-Cloud Icing


This icing condition occurs when supercooled cloud or fog water drop-
lets, 100 μm or less in diameter, collide with a structure. This occurs in
mountainous areas where adiabatic cooling causes saturation of the atmo-
sphere to occur at temperatures below freezing, in free air in supercooled
clouds, and in supercooled fogs that exist in a stable air mass caused by a
strong temperature inversion. Significant accumulations of ice can result.
Large concentrations of supercooled droplets are not common at air tem-
peratures below about 0 °F (−18 °C).
In-cloud icing forms rime or glaze ice with a density between about 10
and 56 pcf (150 and 900 kg/m3), depending on the amount of entrapped
air. If the heat of fusion that is released by the freezing droplets is removed
by convective and evaporative cooling faster than it is released, the drop-
lets freeze on impact. The degree to which the droplets spread as they
collide and freeze governs how much air is incorporated in the accretion,
and thus its density. If the cooling rate is relatively low, not all the colliding
droplets freeze. The resulting ice accretion will be clear or opaque, possibly
with attached icicles.
The collision efficiency of a structure is defined as the fraction of cloud
droplets in the volume swept out by the structure that actually collide with
it. The basic theory of the collision efficiency of smooth, circular cylinders
perpendicular to the flow of droplets carried by a constant wind was devel-
oped by Langmuir and Blodgett (1946). Collision efficiency increases with
wind speed and droplet diameter and decreases as the diameter of the
cylinder increases. For a given wind speed and droplet size, the theory
defines a critical cylinder diameter beyond which accretion will not occur.
This concept of a critical diameter has been confirmed by observation. For-
mulas for calculating collision efficiencies based on an updated numerical
analysis are provided in Finstad and Lozowski (1988).
The amount of ice accreted during in-cloud icing depends on the dura-
tion of the icing condition and the wind speed, as well as on the liquid
Appendix H 169

water content and the size of the droplets in the supercooled clouds or fog.
If, as often occurs, wind speed increases and air temperature decreases
with height aboveground, larger amounts of ice will accrete on higher
structures. The accretion shape depends on the flexibility of the structural
member or component. If it is free to rotate, such as a long guy or a long
span of a single conductor or wire, the ice accretes with a roughly circular
cross section. On more rigid structural members and components, the ice
forms in pennant shapes extending into the wind.

H.1.3 Snow
Sticky snow that falls on a round cross-sectional structural member or
component (such as a wire, cable, conductor, or guy) may deform and/or
slide around it. Due to the shear and tensile strength of the snow resulting
from capillary forces, interparticle freezing (Colbeck and Ackley 1982),
and/or sintering (Kuroiwa 1962), the accreting snow may not fall off the
structural member during this process. Ultimately, the snow forms a cylin-
drical sleeve, even around bundled conductors and wires. The formation
of the snow sleeve is enhanced by torsional rotation of flexible structural
members or components because of the eccentric weight of the snow. The
density of accreted snow ranges from below 5 to 50 pcf (80 to 800 kg/m3)
and may be much higher than the density of the same snowfall on the
ground.
Damaging snow accretions have been observed at surface air tempera-
tures ranging from the low 20 °F up to about 36 °F (−5 °C to 2 °C). Snow
with high moisture content appears to stick more readily than drier snow.
Snow falling at a surface air temperature above 32 °F (0 °C) may accrete
even at wind speeds above 25 mph (10 m/s), producing dense [37 to 50 pcf
(600 to 800 kg/m3)] accretions. Snow with lower moisture content is not as
sticky, blowing off the structure in high winds. These accreted snow densi-
ties are typically between 2.5 and 16 pcf (40 and 250 kg/m3) (Kuroiwa
1965). Dry snow can also accrete on structures (Gland and Admirat 1986).
The cohesive strength of the dry snow is initially supplied by the interlock-
ing of the flakes, and ultimately by sintering, as molecular diffusion
increases the bond area between adjacent snowflakes. These dry snow
accretions appear to form only in very low winds and have densities esti-
mated at between 5 and 10 pcf (80 and 150 kg/m3) (Sakamoto et al. 1990,
Peabody 1993).

H.1.4 Hoarfrost
Hoarfrost is an accumulation of ice crystals formed by direct deposition of
water vapor from the air onto a structure. Because it forms when air with a
dew point below freezing is brought to saturation by cooling, hoarfrost is often
170 Electrical Transmission Line Structural Loading

found early in the morning after a clear, cold night. It is feathery in appearance
and typically accretes up to about an inch (25 mm) in thickness with very little
weight. Hoarfrost does not constitute a significant loading problem.

H.2 LOADING IMBALANCES

Unbalanced loads from in-cloud icing may be significant (White 1999).


Because the rime density and thickness increase with wind speed, signifi-
cant differences in ice loading can occur from one span to the next where
the transmission line crosses a ridge, hill, or escarpment. This can result in
a severe loading imbalance on the line, particularly if adjacent span lengths
are significantly different. When a transmission line is to be located in a
region where in-cloud icing occurs, the engineer would benefit from con-
sulting a meteorologist to determine the severity and extent of the ice
loads. With this information, the engineer can either relocate the line to
reduce the exposure or identify line sections with the greatest risk for
in-cloud icing and adjust designs accordingly.
Snow accretions may shed from wires in the process of formation, before
forming a cylindrical sleeve around the wire. Low-density snow accretions
formed in light winds may shed when the wind speed increases. When
snow sheds from some of the spans, the still-loaded spans will pull slack
from the unloaded spans. This can cause significant increases in the sag of
the wire in the loaded spans. Such events can create clearance violations,
especially if there is deep snow on the ground, even though the associated
unbalanced loads are typically small.
Variations in ice loading during precipitation icing are typically gradual
along the length of a transmission line. Therefore, unequal icing of adjacent
spans is not significant.
Unbalanced longitudinal loadings associated with ice dropping or
unequal ice formation on adjacent spans depend on the relationships
between available slack, insulator lengths, and other factors. Suggestions
for the determination of unbalanced ice loads can be found in IEC 60826
(IEC 2003b) and the various national options of EN 50341 (CENELEC 2012).

H.3 ICE ACCRETION DATA AND MODELING

There are very little data in North America on equivalent uniform ice
thicknesses from natural ice accretions on overhead lines. Therefore, ice
loading studies often rely on mathematical models based on the physics of
the various types of icing and on meteorological data (i.e., precipitation
amount and type, temperature, wind speed) that are required as input to
these models. Results from an ice accretion analysis typically give calcu-
Appendix H 171

lated ice thicknesses for past storms in which freezing precipitation has
occurred. An extreme value analysis can then be applied to determine tMRI.
Wind speeds during and after periods of freezing precipitation can also be
extracted from the meteorological data and analyzed to determine the
wind speed to apply concurrently with tMRI.
There are a number of ice accretion models available that use weather
data to determine accreted ice loads, including the conservative Simple
model (Jones 1998), similar to the Goodwin model (Goodwin et al. 1983),
the US Army Cold Regions Research and Engineering Laboratory (CRREL)
model (Jones 1996), the Makkonen model (Makkonen 1996), the Meteoro-
logical Research Institute (MRI) model (MRI 1977), and the Chaîné model
(Chaîné and Castonguay 1974). The following comments provide informa-
tion on the above mentioned models:

• The Simple model determines the ice thickness, t, from the amount
of freezing rain and the wind speed. t does not depend on the air
temperature because it is assumed that all the available precipita-
tion freezes, and t also does not depend on the wire diameter.
• The CRREL model is less conservative than the Simple model,
using a heat-balance calculation to determine how much of the
impinging precipitation freezes directly to the wire and how much
of the runoff water freezes as icicles. It calculates smaller ice loads
than the Simple model when the air temperature is near freezing
and wind speeds are relatively low; however, water that does not
freeze immediately may freeze as icicles as it drips off the wire. The
CRREL model requires the user to specify the diameter of the wire
on which the accretion of ice is to be modeled. However, this
model, like the Meteorological Research Institute and Makkonen
models, shows very little dependence of ice thickness on wire
diameter.
• The Meteorological Research Institute model tends to determine
smaller ice loads than the CRREL model because water that does
not freeze immediately is ignored, rather than being allowed to
freeze to form icicles. However, in using that model or the Goodwin
model, the user is required to specify the fall speed of the rain
drops and the model results depend significantly on the speed that
is chosen. The Meteorological Research Institute model also deter-
mines accreted snow loads and in-cloud icing loads; however,
many of the significant parameters, including droplet size and
liquid water content of the supercooled clouds, rime accretion
density, and snow sticking fraction and snow accretion density,
must be chosen by the user.
• The Makkonen model for ice accretion in freezing rain tends to be
almost as conservative as the Simple model, primarily because it
172 Electrical Transmission Line Structural Loading

assumes that a significant portion of the water that does not freeze
immediately is incorporated in the accretion. Thus, there is rela-
tively little water available to freeze as icicles.
• The Chaîné model is based on wind tunnel tests that were done by
Stallabrass and Hearty (1967) to investigate sea-spray icing. A number
of assumptions and extrapolations are made to mold these data into a
formulation for freezing rain, and the results indicate a significant
variation of uniform radial ice thickness with wire diameter.
There have been some attempts at model validation. Felin (1988) com-
pared measured maximum ice thicknesses on cylinders of Hydro Quebec’s
Passive Ice Meters (PIM) with Meteorological Research Institute model
results, assuming a drop fall speed of 9 mph (4.1 m/s). Yip and Mitten (1991)
compared 61 PIM measurements with Chaîné, Makkonen, Meteorological
Research Institute, and Goodwin model results using data at nearby weather
stations. Yip (1995) used annual maximum ice thickness data from 235 PIM
sites from 1974 to 1990 and compared the factored ice thicknesses to annual
maxima from the Chaîné model. Jones (1998) compared the measured ice
load on a horizontal cylinder in a single freezing rainstorm with Chaîné,
Meteorological Research Institute, Makkonen, Simple, and CRREL model ice
loads using co-located weather data. Newfoundland and Labrador Hydro
et al. (CEA 1998) reported on the results of a 4-year Canadian Electrical Asso-
ciation (CEA) study comparing ice loads on three test spans with ice loads
determined from the Chaîné, Makkonen, and Meteorological Research Insti-
tute models using weather data measured at the test spans in 22 storm
events. In all these comparisons, the ice accretion models as well as the user
interface between the weather data and the model and the assumptions
made in determining the equivalent uniform radial ice thickness from the
ice measurements were tested.
An alternative approach to using meteorological data and ice accretion
models is to establish ice and wind measurement stations at several loca-
tions in the service area of the utility. The uniform radial thickness, t, can
be determined from the typical cross-sectional area, Ai, of the ice accretion
on a wire of diameter d such that
0.5
d  d2 A 
t= - +  + i  (H-1)
2  4 p 

or from the mass, mi, of a typical ice sample of length L such that
0.5
d  d2 m 
t= - +  + i  (H-2)
2  4 πρ L 
i

where ρi is the density of the ice.


Appendix H 173

In determining ice thicknesses for transmission lines from such data, the
height above ground and orientation of the ice samples to the wind must
be considered. With a sufficiently long period of record and a representa-
tive geographic distribution of these stations, extreme ice loads and concur-
rent wind speeds can be determined.

H.4 EXTREME ICE THICKNESS FROM FREEZING RAIN AND


CONCURRENT WIND SPEEDS

The map of 100-year MRI ice thicknesses from freezing precipitation


with concurrent wind speeds (Figures 2-19 through 2-23 in Chapter 2) was
developed from the same data as the 50-year MRI map in ASCE 7-10
(2010b) and the 500-year MRI map in ASCE 7-16 (2017).

H.4.1 Continental United States and Alaska


Historical weather data from 500 National Weather Service (NWS), military,
Federal Aviation Administration (FAA), and Environment Canada weather sta-
tions were used with the CRREL and Simple models to estimate glaze ice loads
in past freezing rainstorms on wires 33 ft (10 m) above ground, at an orientation
perpendicular to the wind. The station locations are shown in Figure H-1 for
the continental United States and in Figure 2-23 for Alaska. The period of
record of the meteorological data at any station is typically 20 to 50 years.
Accreted ice was assumed to remain on the wire until after freezing rain
ceases and the air temperature increases to at least 33 °F (0.6 °C). The maxi-
mum ice thickness and the maximum wind-on-ice load were determined for
each storm. Severe storms, such as those with significant ice or wind-on-ice
loads at one or more weather stations, were researched in Storm Data (NOAA,
1959–present; a monthly publication that describes damage from storms of
all sorts throughout the United States), newspapers, and utility reports to
obtain corroborating qualitative information on the extent of damage from
the storm. Very little corroborating information was obtained about damag-
ing freezing rainstorms in Alaska, perhaps because of the low population
density and relatively sparse newspaper coverage in the state. Extreme ice
thicknesses were then determined using the peaks-over-threshold method
and the generalized Pareto distribution (Abild et al. 1992, Hoskings and
Wallis 1987, Wang 1991). To reduce sampling error, weather stations were
grouped into superstations (Peterka 1992) based on the incidence of severe
storms, the frequency of freezing rainstorms, latitude, proximity to large
bodies of water, elevation, and terrain. A few stations that were judged to
have unique freezing rain climatology were not incorporated in supersta-
tions. Concurrent wind-on-ice speeds were back-calculated from the extreme
wind-on-ice load and the extreme ice thickness. In calculating wind-on-ice
174 Electrical Transmission Line Structural Loading

loads, engineers should keep in mind that the actual projected area of a glaze
ice accretion may be significantly larger than that obtained by assuming a
uniform ice thickness. Thus, the assumption of a force coefficient of 1.0 for
an ice-covered wire will not be conservative.
Figures 2-19 through 2-23 represent the most consistent and best available
nationwide maps for design ice loads. The icing model used to produce the
map has not, however, been verified with a large set of co-located measure-
ments of meteorological data and ice thicknesses. Furthermore, the weather
stations used to develop this map are almost all located at airports. Structures
in more exposed locations at higher elevations, or in valleys or gorges (for
example, Signal and Lookout Mountains in Tennessee, the Pontotoc Ridge
and the edge of the Yazoo Basin in Mississippi, the Shenandoah Valley and
Poor Mountain in Virginia, Mount Washington in New Hampshire, and Buf-
falo Ridge in Minnesota and South Dakota) may be subject to larger ice thick-
nesses and higher concurrent wind speeds. On the other hand, structures in
more sheltered locations (for example, along the north shore of Lake Superior
within 300 vertical feet of the lake) may be subject to smaller ice thicknesses
and lower concurrent wind speeds. Loads from accreted snow or in-cloud
icing may be more severe than those from freezing rain. In particular, in-cloud
icing, possibly combined with freezing drizzle, appears to be the most
significant icing process in eastern Colorado and New Mexico.

Figure H-1. Locations of weather stations used in preparation of Figures 2-19


through 2-22.
Source: ASCE 7-10 (2010b).
Appendix H 175

H.4.2 Special Icing Regions


Special icing regions are identified in Figures 2-19 through 2-23. As
described above, freezing rain occurs only under special conditions with a
cold, relatively thin surface air layer and a layer of warm, moist air aloft.
Thus, severe freezing rainstorms at high elevations in mountainous terrain
will typically not occur in the same weather systems that cause severe
freezing rainstorms at the nearest airport weather station. Furthermore, in
these regions ice thicknesses and wind-on-ice loads may vary significantly
over short distances because of variations in elevation, topography, and
exposure. In these mountainous regions, the values given in Figures 2-19
through 2-23 should be adjusted based on local historical records and expe-
rience to account for possibly higher ice loads from both freezing rain and
in-cloud icing.

H.5 EXTREME LOADS FROM IN-CLOUD ICING AND STICKY SNOW

Information to produce maps similar to Figures 2-19 through 2-23 for


in-cloud icing and snow accretions is not currently available.

H.5.1 In-Cloud Icing


In-cloud icing may cause significant loadings on transmission lines in
both mountainous regions and level terrain. In the western United States,
in-cloud icing occurs very frequently on exposed ridges and slopes in the
mountains. Above the mean freezing level, heavy deposits can form during
the numerous storms that strike the region in winter. Steep cliff faces and
any exposed structures or obstacles to the wind can become covered with
thick coatings of ice. Although in-cloud icing does not commonly occur
below elevations of about 3,000 ft (915 m), it does occasionally occur when
freezing fog fills the basin regions of eastern Washington and Oregon dur-
ing periods of strong wintertime temperature inversions. In the eastern
plains of Colorado in February 1978, severe rime ice loads were caused by
an upslope fog with winds of 10 to 15 mph (4 to 7 m/s). In Arizona, New
Mexico, and the panhandles of Texas and Oklahoma, the US Forest Service
specifies ice loads due to in-cloud icing for structures constructed at specific
mountaintop sites (USFS 1994). In-cloud icing also occurs in the eastern
United States, primarily on higher peaks in the Appalachian Mountains.
On Mount Washington in New Hampshire [6,280 ft (1,910 m)], the highest
peak in the northeastern United States, in-cloud icing occurs about 50% of
the time from November through April, with icing episodes typically last-
ing less than a day and the temperature remaining below freezing between
episodes. Typical liquid water contents range from 0.3 to 0.6 g/m3 and
176 Electrical Transmission Line Structural Loading

typical wind speeds during icing range from 31 to 62 mph (14 to 28 m/s),
with wind speeds greater than 90 mph (40 m/s) occurring 2% of the time.
On the more numerous 4,000 ft (1,200 m) mountain summits, in-cloud icing
is less severe because the peaks are not exposed to supercooled clouds as
frequently and wind speeds are lower. In-cloud icing loads are sensitive to
terrain exposure and to the direction of the flow of moisture-laden clouds.
Large differences in ice thickness can occur over a few hundred feet dis-
tance and can cause severe load unbalances. Advice from a meteorologist
familiar with the area is particularly valuable in these circumstances.

H.5.2 Snow
Snow accretions can occur anywhere that snow falls, even in regions that
may experience only one or two snowstorms a year. In some regions,
extreme accreted snow loads are greater than ice loads from freezing rain
or drizzle. A heavy, wet snow storm on March 29, 1976, caused $15 million
in damage to the electric transmission and distribution system of Nebraska
Public Power District (1976). Mozer and West (1983) reported a transmis-
sion line failure on December 2, 1974, near Lonaconing, Maryland, due to
heavy, wet snow of 5 in. (127 mm) radial thickness on the wires with an
estimated density of 19 pcf (304 kg/m3). Goodwin et al. (1983) reported
measurements of snow accretions on wires in Pennsylvania with an
approximate radial thickness of 4 in. (102 mm). The meteorological condi-
tions along a transmission line that failed under vertical load in the Front
Range of Colorado were analyzed after the failure. The study indicated that
the failure was caused by a 1.7 inch (43 mm) radial thickness, 30 pcf (480
kg/m3) wet snow accretion with a 42 mph (19 m/s) wind. The return
period for this snow load was estimated to be 25 years (McCormick and
Pohlman 1993). In the winters of 1994–1995 and 1996–1997, Golden Valley
Electric Association in Fairbanks, Alaska, made 27 field measurements of
the radial thickness and density of dry snow accretions. Densities ranged
from 1.4 to 8 pcf (22 to 128 kg/m3) and radial thicknesses were up to 4.4 in.
(112 mm). The heaviest were equivalent in weight to a 1 in. (25 mm) uni-
form radial thickness of glaze ice (Golden Valley Electric Association,
unpublished data, 1997). Finstad et al. (2009) describes the modeling of
sticky snow loads in Alberta, Canada, using weather data.

H.6 OTHER SOURCES OF INFORMATION

Bennett (1959) presents the geographical distribution of the occurrence


of ice on utility wires from data compiled by various railroad, electric
power, and telephone associations covering the 9-year period from the
winter of 1928–1929 through the winter of 1936–1937. The data include
Appendix H 177

measurements of all forms of ice accretion on wires, including glaze ice,


rime ice, and accreted snow, but does not differentiate between them. Ice
thicknesses were measured on wires of various diameters, heights aboveg-
round, and exposures. No standardized technique was used in measuring
the thickness. The maximum ice thickness observed during the 9-year
period in each of 975 squares, 60 miles (97 km) on a side, in a grid covering
the contiguous United States was reported. In every state except Florida,
measurements of accretions, with unknown densities, of approximately
1 inch radial thickness were reported. The map shows measurements as
high as 2 inches (51 mm) in the Northeast, Southeast, and South; 1.75 inches
(44 mm) in the Midwest; 2.4 inches (61 mm) in the High Plains; and 3 inches
(76 mm) in the West. Information on the geographical distribution of the
number of storms in this 9-year period with ice accretions greater than
specified thicknesses is also included in the Bennett report.
Tattelman and Gringorten (1973) reviewed ice load data, storm descrip-
tions, and damage estimates in several meteorological publications to esti-
mate maximum ice thicknesses with a 50-year return period in each of
seven regions in the United States.
In Storm Data, storms are sorted by state within each month. The com-
pilation of this qualitative information on storms causing damaging ice
accretions in a particular region can be used to estimate the severity of ice
and wind-on-ice loads. The Electric Power Research Institute (EPRI) has
compiled a database of icing storms from the reports in Storm Data. Dam-
age severity maps have also been prepared (Shan and Marr 1996).
Robbins and Cortinas (1996) and Bernstein and Brown (1997) provide
information on freezing rain climatology for the 48 contiguous states based
on meteorological data. For Alaska, available information indicates that
moderate to severe ice loads of all types can be expected. The measure-
ments made by Golden Valley Electric Association are consistent in mag-
nitude with visual observations across a broad area of central Alaska
(Peabody 1993). Several meteorological studies using ice accretion models
to determine ice loads have been conducted for high-voltage transmission
lines in Alaska (Richmond 1985, 1991, 1992; Gouze and Richmond 1982a,
b; Peterka et al. 1996). Glaze ice accretions for a 50-year return period range
from 0.25 to 1.5 inches (6 to 38 mm) radial thickness, snow from 1.0 to
5.5 inches (25 to 140 mm) radial thickness, and rime from 0.5 to 6.0 inches
(12 to 150 mm) radial thickness. The assumed accretion densities were
glaze 57 pcf (910 kg/m3), snow 5 to 31 pcf (80 to 500 kg/m3), and rime
25 pcf (400 kg/m3). These ice thicknesses are valid only for the particular
regions studied and are highly dependent on the elevation and local terrain
features. Large accretions of snow have been observed in most areas of
Alaska that have overhead lines.
In areas where little information on ice loads is available, it is recom-
mended that a meteorologist familiar with atmospheric icing be consulted.
178 Electrical Transmission Line Structural Loading

It should be noted that taller structures may accrete more ice because of
higher winds and colder temperatures aloft, and that the influences of
elevation, complex relief, proximity to water, and potential for unbalanced
loading are significant.

H.7 CURRENT PRACTICE

A 1979 survey of design practices for transmission line loadings (ASCE


1982) obtained responses from 130 utilities operating 290,000 miles
(470,000 km) of high-voltage transmission lines. Fifty-eight of these utilities
specifically indicated “heavy icing areas” as one reason for special loadings
in excess of NESC requirements. Design ice loads on conductors ranged
from no ice (primarily in portions of the southern United States), up to a 2
or 2.25 inch (50 or 57 mm) radial thickness of glaze ice in some states.
Radial glaze ice thicknesses between 1.25 and 1.75 inches (32 and 45 mm)
are commonly used. Most of the responding utilities design for heavy ice
on the wires with no wind and less ice with wind. Few utilities consider
ice on the supporting structures in design.
The National Electrical Safety Code (IEEE 2017) defines four geographical
loading districts that specify combined ice and wind loads. A discussion is
provided in the handbook published with the 3rd edition of the NESC
(National Bureau of Standards 1920):

The assumed ice loadings have been chosen after careful consider-
ation of data obtained from the U.S. Weather Bureau, from electric
companies, and from engineers. The values chosen do not represent
the most severe cases recorded, but do represent conditions that occur
more or less frequently. Ice loading of 1/2 inch is frequently exceeded,
particularly near the northern and eastern borders of the U.S., and on
occasions ice has been known to collect to a thickness of 1.5 inches
and even more.

In addition to NESC loading districts (Heavy, Medium, Light, and Warm


Island), the NESC has adopted the 50-year glaze ice map in the previous
editions of this manual (ASCE 1991, 2010a) to establish ice loading criteria.
APPENDIX I
SUPPLEMENTAL INFORMATION REGARDING
LONGITUDINAL LOADS

I.1 INTRODUCTION

Longitudinal loads are complex and may be created through many dif-
ferent events including differential intact wire tensions, broken wires, and
construction loads. When considering longitudinal loads, dead-end struc-
tures are, by definition, capable of supporting full wire tension loads with
all wires removed in one longitudinal direction. Strain, angle, and tangent
structures are typically designed for longitudinal loads much less than full
(one-side only) wire loads. This appendix will focus primarily on longitu-
dinal loading criteria, calculation methods, and failure containment
approaches for tangent, angle, and strain structures. A discussion on trans-
verse cascades is also included.
Longitudinal loads may exceed those imposed by the intact wire system.
The potential for extreme longitudinal loads necessitates that utility own-
ers consider including longitudinal loads as part of their standard struc-
tural loading criteria. Due to the diversity in regions, reliability needs, and
structure types, the approach to longitudinal loading criteria taken by util-
ity owners varies greatly. Longitudinal loading criteria should be collab-
oratively developed by key stakeholders that have an influence on the
reliability of a transmission line. The concepts in this manual have been
published in the industry and should be considered for an owner’s longi-
tudinal loading criteria. It should be noted that longitudinal loading is a
very complex issue and the calculation and failure containment approaches
are not limited to those presented in this manual.

179
180 Electrical Transmission Line Structural Loading

I.2 LONGITUDINAL LOADS ON INTACT WIRE SYSTEMS

Transmission structures are typically designed to resist the longitudinal


imbalances caused by differential tensions in adjacent spans. These dif-
ferential tensions can be caused by significantly differing adjacent span
lengths, topography, and by differing environmental conditions on adja-
cent spans. One such example would be unequal ice loading. This can
result from unequal ice deposits on adjacent spans, or from one span shed-
ding its ice before the ice is shed on the adjacent span.
When calculating the longitudinal loading for a structure, it is important
to consider the ability of the wire supports (i.e., insulators) to swing longi-
tudinally, thereby reducing the longitudinal loading on the structure. A
typical conductor suspension assembly can move longitudinally, which
helps reduce the tension imbalance in the conductors and can reduce the
longitudinal loading on the structure. The shorter the suspension assembly,
the more the longitudinal loading will be transferred to the structure. Post
insulators, depending on their length, end support conditions, and the
material used, can have varying degrees of flexibility. Shield wire attach-
ments are generally considered rigid connections, much like strain
insulators.
In addition to the flexibility of the wire attachments, the longitudinal
flexibility of the supporting structure itself can affect the longitudinal load-
ing experienced by the structure. As the structure deflects due to the load
imbalance, the load imbalance is often reduced. Disregarding the flexibility
of the structural system (structure and wire attachments) can result in
unnecessarily conservative structure design.

I.2.1 Suspension Supports


Unequal wind or ice loads on adjacent spans and conductor temperature
variation on unequal adjacent span lengths can result in differential wire
tensions. On suspension structures, the resultant structural load due to
differential wire tensions is usually reduced by the swing of the suspension
insulator strings (Cluts and Angelos 1977). The longitudinal loads trans-
mitted to the structures by the inclined suspension strings rarely exceed
10% to 20% of the conductor bare wire tension, except in hilly or mountain-
ous terrain where in-cloud icing is a hazard. Suggestions for the determina-
tion of unbalanced ice loads can be found in IEC 60286 (2003b), EPRI
EL-643 (1978), and EN 50341 (CENELEC 2001).
Many lines exist with unequal spans. A single suspension supported
short span may not present a problem because it takes very little longitu-
dinal movement at the support to equalize the tensions. However, a span
that is significantly longer than the others will attempt to keep the tension
balanced at the supports as the wire contracts or elongates. In this case,
Appendix I 181

more longitudinal movement at the support than is required to relieve the


tension imbalance may not be available and the imbalance longitudinal
load is transferred to the structure. Similarly, one needs to exercise caution
where there are several long spans in succession followed by several short
spans. The longitudinal imbalance or movement can be quite large at the
transitions. Designers may place strain or dead-end structures at the transi-
tions between span lengths.

I.2.1.1 Impact of Slack in Adjacent Spans­  In the case of in-cloud icing,


the longitudinal loads can be significant because ice deposits can vary
greatly from span to span (see Chapter 2, Section 2.3.6 and Appendix H).
Unloaded adjacent spans with significant slack permit the insulator to
swing sufficiently to turn the suspension assembly into a strain support
that is likely to transfer nearly all the differential tension to the structure.
In this context, slack is defined as the difference between the actual wire
length and the straight-line distance between the attachment points.
Problems have been observed in areas where the slack difference of
adjacent spans exceeds twice the length of the insulator strings.
Assuming level spans and using the parabolic approximation of the
catenary, the following relationships may be used:
wS 2
sag @ (I-1)
8TH

sag 2 S3
slack ≅ 8 = 2 (I-2)
3S  TH 

24  
 w 

where
w = Wire unit weight,
S = Straight-line span length, and
TH = Horizontal component of tension.

When the conductor within a ruling span section is not longitudinally


restrained (generally true at suspension insulators), slack is proportional
to the cube of the ruling span. For example, using a TH/w ratio of 5,000 ft,
the slack in an 800 ft span equals 0.85 ft. Using the cube relationship, the
slack in a span double the length (1,600 ft) would be 8 × 0.85, which equals
6.8 ft.

I.2.1.2 Shield Wire Supports  Shield wire supports often have short
suspension linkages that provide insignificant reduction of unbalanced
wire loads. Differential tensions may develop at shield wire supports
182 Electrical Transmission Line Structural Loading

during significant changes (low to high) in ambient temperatures.


Generally, designers will specify an appropriate longitudinal load to
account for these changes in temperature.
If the line is located in an in-cloud icing area, the shield wire support of
suspension structures may pose the highest risk for structure failure
because the differential shield wire tensions could be significantly higher
than the differential conductor tensions produced by the same weather
conditions. Designers have utilized several methods to reduce the differ-
ential shield wire tensions caused by in-cloud icing. Examples include lon-
ger suspension links to provide more flexibility to reduce tension imbalance;
slip or release clamps to limit the maximum load acting on a support point;
and, in some cases, shield wire supports have been designed to act as fuses
to collapse at defined loads, thereby preventing more serious damage.
Some utilities have removed the shield wires from lines located in areas
likely to experience in-cloud icing. However, if removing the shield wire,
the designer should address the potential impact on grounding and light-
ning protection.

I.2.2 Strain Supports


Strain supports by definition are designed to resist differential tensions
(longitudinal loads) from adjacent spans, but they may be designed for less
than full dead-end capability.

I.3 LONGITUDINAL LOAD CALCULATIONS


If the failure of a component, such as a broken wire, could cause a cas-
cading failure of successive tangent structures not designed to resist wire
tension loads, failure-related load criteria should be considered to mini-
mize the extent of the damage and the time required to restore service
(EPRI 1979). Failure-related load criteria, such as the broken wire load
(BWL), have been used successfully to mitigate the effects of severe dif-
ferential wire tensions and to minimize the extent of a failure. Based on
experience (Ostendorp 1997), the breakage of conductors, shield wires,
components, and line cascades are serious problems. The potential for a
cascade exists when sufficient slack is introduced into a span so that the
unbalanced longitudinal load at the adjacent structure is significant enough
that it could fail that structure. As the second structure fails to resist the
residual load, it allows the wire to move on to the next structure and repeat
the sequence, resulting in a cascading failure. To stop the cascading, it is
necessary to limit slack transfer.
While the dynamic loads immediately after a phase or wire break could
cause instantaneous loads to be higher than the static intact condition, it is
Appendix I 183

typical industry practice not to include these dynamic loads in the design
of structures. Dynamic loads are localized, impact only the adjacent three
structures or less, and quickly dissipate.
Most structure types can be designed to provide some longitudinal
strength resulting in increased resistance to cascading failure for a small
increase in initial cost. Two common, simple methods are often used to
estimate an unbalanced longitudinal load: the Residual Static Load (RSL)
Method and the EPRI Method.
It should be noted that the unbalanced longitudinal loads determined
using RSL and EPRI methods constitute the minimum required “static”
loads to be resisted by the structures to avoid cascading failures. Note: The
RSL factors do not consider dynamic effects.) The calculated unbalanced
longitudinal loads act on the support structure in the direction away from
the initiating failure event and should be considered to act concurrently
with the effects of any permanently applied load imbalance.

I.3.1 Residual Static Load Method

The RSL Method determines a longitudinal load factor that is applied


to the wire tension in order to determine the residual static load (RSL) on
structures after a wire failure and after all dynamic effects from the wire
break have diminished. The reduction in the load magnitude resulting
from the insulator swing and support deflection may be considered in the
calculation of the RSL. Computer programs (EPRI 1983, Mozer et al. 1977,
Peyrot 1985) and design charts (Comellini and Manuzio 1968, EPRI 1978)
can be used to assist in the calculation of the RSL values. RSL values
derived from the Comellini and Manuzio charts are based on insulator
string length and span length and will provide values of approximately
60% to 70% of everyday tensions.
Figure I-1 provides RSL longitudinal load factors as a function of the
span/sag ratio and the span/insulator ratio. The calculation assumes rigid
supports (i.e., the potential benefiting effects of the flexibility of the sup-
ports are neglected) and 10 equal-length spans between the wire break and
the next dead end. The span/insulator ratio is the ratio of the average span
length within a given tension section to the average effective insulator
length (i.e., the insulator length free to swing longitudinally).
The RSL is calculated for the bare wire (no ice or wind) loading condi-
tion at an average temperature. The RSLs are applied to one of the conduc-
tor support points or to one (or both) shield wire support point(s) on a
structure. RSLs are applied in only one longitudinal direction, along with
50% or more of the intact wire vertical load. The other support points on
the structure will be considered in the intact condition.
184 Electrical Transmission Line Structural Loading

Figure I-1. RSL Method longitudinal load factor.

I.3.2 Electric Power Research Institute Method


The Electric Power Research Institute (EPRI) developed another meth-
odology for calculation of unbalanced longitudinal loads. This method was
developed from research completed by EPRI, and calculates loads as a
function of the horizontal wire tension, the span/sag ratio, the span/insu-
lator ratio, and the support flexibility (Ostendorp 1997). Although this
method can also approximately predict the impact load on the structures
adjacent to the initial failure, due to the complexity of those calculations,
that portion of the EPRI method is not presented here.
Figure I-2 provides longitudinal load factors as a function of the span/
sag ratio and the stiffness of the support structures. Wire tensions multi-
plied by the longitudinal load factors provide approximate design loads
that include dynamic effects, structural stiffness, and insulator lengths.
The span/sag ratio is the ratio of the average span length within a given
tension section to the sag of the average span for a given conductor or
shield wire tension. Longitudinal load factors are provided for “rigid”
structures such as guyed or latticed structures of high stiffness, as well as
for “flexible” structures such as single poles capable of enduring large
elastic deformations.
Appendix I 185

Figure I-2. EPRI Method longitudinal load factor.

I.4 FAILURE CONTAINMENT APPROACHES

Transmission lines may be exposed to severe wind and ice loads, vehicu-
lar impact, and other extreme events that may result in structural loading
exceeding the criteria for which the transmission structures were designed.
Additionally, if a structural or hardware failure occurs, longitudinal load-
ing may exceed design structural loading on several adjacent structures.
When longitudinal loading exceeds design structural loading, infrequent
failures of a few structures or components due to these extreme events is
a generally accepted practice among utilities and utility maintenance prac-
tices should be planned accordingly. However, it is recommended that the
designer of the line consider developing structural loading criteria that are
coordinated with the maintenance practices of the utility. The structural
loading criteria should also provide the utility with the desired transmis-
sion line performance under extreme loading so that transmission line fail-
ures are contained as desired.
When selecting cascade failure mitigation methods, the designer should
consider several factors:
• The inherent longitudinal strength of the structure types being
used: Square-based latticed towers are inherently stronger in the
longitudinal direction than H-frame structures. Including a modest
186 Electrical Transmission Line Structural Loading

longitudinal design load may not affect the cost of the latticed
towers, while it may be too costly or not feasible for a self-sup-
ported H-frame structure.
• Criticality of the transmission line: Lines that do not have any
redundancy or have such high demand that an extended,
unplanned outage would cause issues elsewhere in the system are
often considered critical assets. As such, increased up-front spend-
ing to significantly limit the extent of any potential failure may be
warranted.
• Degree of difficulty in restoring the line: Some lines are more
difficult to restore than others. There may be significant topography
or environmentally sensitive areas that make access difficult. Some
structures on the line may be unique and difficult to replace, such
as extra-tall structures at a crossing. The subsurface conditions may
make locating or constructing foundations difficult (e.g., very hard
rock, karstic formations, significant below-grade infrastructure).
• Cost and availability of replacement materials: If a line can be easily
restored with materials that are readily available, it may be less
critical to limit the extent of a failure. However, if a line could only
be restored using expensive or long-lead materials, then it might
make more sense to pursue increased failure containment
measures.
• Construction cost differential for each method: Some structures
have inherent longitudinal strength and as such can be designed for
failure containment loads with little or no increase in costs. In cases
where designing each structure for failure containment loads
would not be economical, it may make more sense to install peri-
odic stop structures.

Successful failure containment may be achieved by providing sufficient


longitudinal strength (1) on all structures, or (2) on failure containment
structures inserted at regular intervals. Angle structures may be used as
failure containment structures if their longitudinal strength is sufficient to
resist the unbalanced loads and to arrest a cascading failure.

I.4.1 Failure Containment Philosophy


Initial transmission line failures may be caused by several different
types of events such as train derailment, a major tornado, a low-flying
aircraft, or a severe ice storm. An event may bring several structures to the
ground and may be accompanied by component and wire failures. The
failures could create complicated dynamic forces at adjacent structures.
The arduous effort needed to quantify the dynamic energy or impact
Appendix I 187

component at the adjacent structures has directed attention to the security


(or survival) of the second, third, fourth, or fifth structure away from the
initial failure (Thomas 1981, Ostendorp 1997, Kempner 1997).
Depending on the importance of the line, it is generally agreed that if
the second, third, fourth, or fifth structure from the initiating event does
not fail, there will be no cascade and most of the energy released by the
failure will have dissipated. Therefore, the problem of failure containment
may be reduced to the problem of determining the required longitudinal
strength to resist the differential tensions at the second, third, fourth, or
fifth structure, respectively, while allowing the failure of one or more struc-
tures to dissipate the released energy.

I.4.2 Basic Assumptions


It should be noted that any event that permits the creation of excessive
slack is likely to produce longitudinal loads that may lead to a cascading
failure. Longitudinal cascades of high-voltage lines (Frandsen and Juul
1976) have resulted from initial failures other than broken shield wires or
conductors. For example, failure of a heavy angle structure could introduce
excessive conductor slack and longitudinal loads that trigger cascading
failures on both sides of the fallen angle structure.

I.4.3 Failure Containment Approaches


H-frames and narrow-based, rectangular, latticed structures have little
inherent ability to withstand the longitudinal loads of a cascading line.
Additionally, the shield wires attached to these structures with near-rigid
attachments may contribute to or initiate a cascade. It is considered pru-
dent design practice to employ methods to limit the length of a cascade.
For existing lines with limited longitudinal strength, the cost of strengthen-
ing such structures or adding longitudinal guys to tangent supports is
likely to be prohibitive, undesirable, or ineffective. Another option would
be to insert failure containment structures (e.g., stop structures, anchor
structures, anti-cascading structures, full dead-ends) at prescribed intervals
along the line to limit the extent of the damage caused by a component,
structure, or foundation failure.

I.4.3.1 Failure Containment Structures  Design codes, criteria, and philos­


ophies for failure containment differ by regions, transmission line owners,
and system operators. A typical distance interval between failure
containment structures varies but may be as long as 16.1 km. Judgment of
these distances may include the length and importance of the line,
longitudinal strength of the suspension structures, terrain, land use,
188 Electrical Transmission Line Structural Loading

restoration time, emergency stocking levels, cost, right-of-way access, and


proximity to facilities that may be affected, such as railways, interstate
roadways, etc. The decision to install or create anti-cascade or stop towers
at intervals along an existing line requires an awareness of the means by
which a longitudinal cascade is propagated. Failure to appreciate the
mechanics involved may negate the entire effort.
One consideration that should not be overlooked in the placement of
anti-cascading structures is their role in the construction stringing process.
Anti-cascading structures could be located with input from experienced
construction personnel to incorporate stringing staging, segment construc-
tion, and stringing length limitations. This can allow the designer to
include anti-cascading structures at the aforementioned intervals and pro-
vide construction crews more optimal structure configurations for
stringing.
Special resistance structures are typically latticed towers or guyed or
self-supported pole structures that provide a sufficient level of strength to
resist the unbalanced longitudinal loads caused by the failure of compo-
nents, wires, or structures. Special resistance structures have traditionally
included a structure capable of supporting wires only on one side or with
broken wire capability. In general tangent (suspension) structures do not
possess sufficient strength to provide resistance to cascading failures. The
cost associated with developing such capacity in these suspension struc-
tures may be prohibitive.
It is important to note that for high voltage (HV) and extra high voltage
(EHV) lines, the attempt to use a rigid suspension structure as a stop tower
will not succeed even if the tower itself has great longitudinal strength.
Allowing the wire movement to pass on down the line will ensure continu-
ance of the cascade even though the stop towers may remain standing. In
such instances, it is recommended that special resistance structures be pro-
vided at selected intervals along the line to limit the length of a cascading
failure to an acceptable number of structures. A strain-type structure could
stop a cascade if it has sufficient strength to resist the unbalanced loads
(bare wire or iced, as required) and prevent the movement of wire along
the line. A suitably strong strain-type angle tower will serve this purpose.
In new line construction, the frequent need for angle structures may be
accepted as a design alternative to building the anti-cascade strength into
each suspension structure.
On low-voltage H-frames or portal structures, the length of the insulator
strings will approximately equal the available deflection; therefore, appli-
cation of longitudinal storm guys, and possibly the installment of metal
crossarms at appropriate intervals, can be a rational means of removing the
threat of long cascades. Such applications at major highway or railroad
crossings should consider the impact of broken wire loading on the behav-
ior of containment structures. At more general locations, RSL may be
Appendix I 189

specified because the impact is diminished to a residual tension after a


number of structures a distance away from the break.
The alternative of either inserting or converting a suspension-type struc-
ture to perform the anti-cascading duty is attractive but not always pos-
sible if the suspension strings are long, as with high voltage (HV) and extra
high voltage (EHV) lines. It is possible that enough wire movement will be
passed on through the stop tower so the failures continue. The swing of
the insulator string will produce a longitudinal load equal to the vertical
load being supported at the point, multiplied by the tangent of the angle
of swing of the insulator string. This secondary effect must be checked.

I.4.3.2 Install Load Release Mechanisms  Release mechanisms that


reduce the tension or drop the conductor have been developed but are not
widely accepted or used (Peabody 2004). These devices can be divided into
two types: (1) Energy-absorbing mechanisms which reduce impact loading
and add slack into the span, thereby reducing the RSL; and (2) load release
mechanisms which allow the wire to slip through the suspension clamp.
The performance of release mechanisms should be calibrated and
verified in representative tests. It is imperative that the design of the slip
or release mechanism perform adequately under the design climatic and
operational conditions. Regardless of the release device deployed, the
structure and supporting hardware must be able to withstand the RSL that
the release mechanism does not shed.
Some release mechanisms may not be suitable for use in areas where
heavy ice buildups are frequent. Premature release of the device under
unbalanced ice could result in a dangerous and undetected ground clear-
ance violation that may constitute a danger to the public.

I.4.3.3 Limited Structure Failure Design  Transmission structures can be


designed to have a specific component failure mode so that a structural
component such as the arm will have less longitudinal capacity than the
structure body (Bryant 2012). In this case, the longitudinal capacity of the
arm would be designed so that it would fail prior to other structure
components. This would allow the longitudinal load to balance and/or
dissipate while the core structure body remains undamaged. This can lead
to a decrease in recovery time after a failure event by limiting the structure
damage to certain components. Designers should also consider the
durability, mode, and rate of local component failure. In general, plastic
failure is considered to be preferable to shear or tension failure in
mechanical fuses that may fail rapidly and induce a dynamic load into the
system.

I.4.3.4 Failure Containment for Icing Events  In areas where icing events
are frequent, utilities may adopt failure containment loads with iced
190 Electrical Transmission Line Structural Loading

conductors as a design requirement for important lines. Specially reinforced


or guyed structures may be used at regular intervals to resist the extremely
large differential tensions. These same structures may also be capable of
arresting a cascading failure.

I.4.3.5 Failure Containment: Bonneville Power Administration Method 


A system approach can be followed to mitigate the effects of failure-related
unbalanced longitudinal loads on transmission lines. The system approach
(Kempner 1997) uses a “failure containment” philosophy that accepts the
failure of one tower on each side of the initiating event. The longitudinal
loading case assumptions are (1) only one wire or phase is broken at one
time, and (2) the break occurs during an everyday load situation, which is
defined as no ice, no wind, a conductor temperature of 30 °F (−1.1 °C), and
initial sag. The conductor tension obtained under these conditions is
multiplied by an impact factor. Standard suspension towers (0° to 3° line
angle) and “heavy” suspension towers (0° to 6° line angle) have an impact
factor of 1.33. The impact factor for “light” suspension towers (no line
angle) is 0.67.

Suspension Tower Conductor  The load case consists of


• A vertical load at the broken conductor attachment point [i.e., 50%
of the conductor weight and hardware at 30 °F (−1.1 °C)] and the
vertical load at the attachment point of intact conductors (i.e., the
weight of the conductors and hardware).
• A longitudinal load at the support (i.e., bare wire everyday tension
multiplied by the appropriate impact factor).
• A transverse load caused by line angle. Only one phase is assumed
broken for both single- and double-circuit towers. Each conductor
attachment point shall be considered individually. For a double-
circuit tower, this load case shall be repeated with only one circuit
strung.

Strain Dead-End Conductor  The load case consists of


• A vertical load (i.e., weight of the conductor and hardware) at 0 °F
(−17.8 °C).
• A transverse wind load on the tower and wires [i.e., at 40 mph
(18 m/s)] with no ice.
• A longitudinal load equal to 125% of sagging tension. The vertical,
transverse, and longitudinal wire load is multiplied by a 1.5 load
factor. For double-circuit towers, this load case shall be repeated
with only one circuit strung.
Appendix I 191

Shield Wire  The load case consists of


• A vertical load of the iced shield wire (i.e., weight of glaze ice
equivalent to 1.5 times the working load shield wire design ice
thickness at maximum working tension). The equivalent glaze ice
thicknesses are light suspension tower, 0.75; standard suspension
tower, 1.125; and heavy suspension tower, 1.125. The vertical load is
the sum of one-half the equivalent iced wire weight for 1.5 times
the transverse span plus one-half the bare weight of 0.5 times the
transverse span. Additionally, a vertical conductor load equal to the
equivalent ice-coated wire weight is applied to 1.0 times the
transverse span.
• The longitudinal load of the shield wire equals the horizontal
tension and is applied to all shield wire peaks.

I.4.3.6 Percent of Everyday Wire Tension  A design longitudinal load,


historically known as broken wire load (BWL) (ASCE 1991), can also be
used. Experience has shown that flat or horizontal configurations, single-
circuit lines designed with the BWL concept have produced transmission
lines with a sufficient level of longitudinal strength to contain the effects
of broken wires and other comparable failures that may have otherwise
resulted in a cascade.
A horizontal load that is equal to the everyday bare wire tension (EDT)
of the shield wire and is equal to about 70% of the EDT of a conductor,
applied as a single load at any one support point, has been used success-
fully to mitigate the effects of broken wires. It should be noted that fre-
quently occurring heavy ice conditions or stiff, brittle supports may require
a larger longitudinal load.

I.5 TRANSVERSE CASCADES

I.5.1 Characteristics of a Transverse Cascade


Transverse cascades are differentiated from longitudinal cascades in that
the tension, magnitude, and direction of the load of the wire system after
the collapse of the initiating structure failure are predominately in the
transverse direction. Successive structural failures in a transverse cascade
collapse in a generally transverse direction; as such, they may be incor-
rectly considered a failure caused by a broad front wind. A plan view (top
view) of a typical transverse cascade is illustrated in Figure I-3.
Transverse cascades can occur in various ways or from different causes.
Examples include (1) HIWs that overload one structure transversely and
192 Electrical Transmission Line Structural Loading

Figure I-3. Plan view of typical transverse cascade.

its failure successively overloads adjacent structures, and (2) on short spans
of wood poles, a weak or decayed pole may fail due to transverse wind,
which adds wire tension load to the wind load on adjacent structures that
then also fail transversely.
Most transverse cascades are initiated by the impact of a high-intensity
wind (HIW) on the line, with one or two structures brought down. These
small, local failures frequently become transverse cascades of dozens of
structures. These failure scenarios have often been misjudged as multiple
failures caused by a “wall of wind” overcoming all the fallen structures,
when the actual failure mechanism was an initial failure of a single struc-
ture due to HIW and subsequently a transverse cascade. Failure of many
towers from widespread transverse wind is not common except in areas
subject to cyclones, hurricanes, or seaside gales.
An understanding of the loads generated in the wire system after the
transverse collapse of one or two structures from HIWs along with an
awareness of the line systems (wires and structures) that are vulnerable to
these loads can assist the line engineer in providing design options that
will reduce the possibility of these types of failures.

I.5.2 Wire Behavior of a Transverse Cascade


A significant parameter in what follows is the slack, which is the differ-
ence in length between the straight line joining the points of support and
the length of the suspended wire. This exercise uses a parabolic equation
because the added precision of working with catenaries is not required. See
Equations (I-1) and (I-2). In these two equations, sag is a function of span2
and slack is a function of span3.
The ratio of TH/w is generally referred to as the catenary or parabolic
constant. For typical spans with the parabolic constant of tension/unit
weight of 487.7 m can be found in Table I-1.
It may be noted that the conductors are supported on suspension insula-
tor strings permitting restricted longitudinal swing varying with the length
of the string, whereas the shield wires are almost always firmly attached
to the tops of the shield wire peaks of the structures.
Appendix I 193

Table I-1.  Typical Span Characteristics

Span Sag Slack

400 ft (121.9 m) 12.5 ft (3.81 m) 1.04 ft (0.317 m)

800 ft (243.8 m) 50.0 ft (15.24 m) 8.33 ft (2.54 m)

1,200 ft (365.8 m) 112.5 ft (34.29 m) 28.13 ft (8.57 m)

1,600 ft (487.7 m) 200.0 ft (60.96 m) 66.67 ft (20.32 m)

With the transverse failure of a single structure, the added length to the
wire system can be calculated as well as the transverse and longitudinal
loads applied to the adjacent structures. The insulator strings will swing
toward the fallen structure, pulling slack from adjacent spans of conduc-
tors. However, with the resistance offered by the inclined insulator strings,
there will be a great increase in all conductor tensions. These tensions exert
longitudinal forces on these towers as well as significant transverse loads.
Shield wire tensions will increase more rapidly with no relief due to
insulator string swing, and the pulls exerted on the tops of the shield wire
peaks will be limited only by the slip strength of the clamps or the fusing
capacity of the shield wire peak itself.
These loads can overwhelm the adjacent two structures, leading to a
compression buckling of the mast or nearest corner leg of a latticed struc-
ture. However, as the structure starts to fall, the inward tensions start to
relax while the tensions back to the next set of adjacent structures will
increase. The falling structures will therefore describe an arc in falling,
pulled first toward the failed structure but then away from it. Crossarms
will strike the ground slightly away from the trigger structure, sometimes
as much as 3 ft (1 m).
This pattern of structures falling slightly away from the trigger structure
can be readily discerned on site if the investigator is aware of the phenom-
enon. If the structures on the ground almost “point” back toward the trig-
ger tower and there is further evidence of the failed and outwardly splayed
corner legs of a latticed structure, the sequence of events can be confirmed.

I.5.3 Conditions Leading to Transverse Cascading


By examining the parameters that will create the greatest diagonal pulls
on adjacent structures, on the one hand it can be noted that
• Short spans contain little slack to relieve the high tensions pro-
duced by the falling structure. Short spans also create the greatest
tension increases after the failure of one structure.
194 Electrical Transmission Line Structural Loading

• Tall structures (such as double-circuit vertical configurations), in


falling transversely, lead to large increases in wire loads of upper
conductor phases and of the shield wires.
• The short insulator strings of low-voltage lines restrict movement
of slack from adjacent spans.
On the other hand, EHV lines are inherently safer with regard to trans-
verse cascading for several reasons:
• Longer insulator strings permit greater equalization or reduction of
conductor tensions.
• Strength requirements for carrying the bundled conductors of an
EHV line minimize the influence of the shield wire system that
usually is similar to that used for lower voltages.
• The longer spans usually associated with EHV also contain larger
amounts of slack and do not tighten as quickly when one structure
falls.
It should be noted that the reduced influence of the shield wire system
on EHV lines may be threatened by the increasing trend of replacing con-
ventional small steel or aluminum-clad steel wire stranding with much
larger, heavier, and stronger optical ground wire (OPGW). Replacement
with OPGW may require a corresponding strengthening of the clamping
and the shield wire peaks themselves.
The shield wire system can, and in most cases does, contribute a major
part of the cascade-inducing forces because it is the highest part of the wire
system, and the direct clamp system permits no equalization or reduction
of tension. A stronger shield wire system, comprised of wires, clamps, and
shield wire peaks, increases the potential for a transverse cascade.
APPENDIX J
INVESTIGATION OF TRANSMISSION
LINE FAILURES

J.1 INTRODUCTION

Line failures provide a unique and highly valuable opportunity to


increase our understanding of transmission line behavior. Not all damage
or failures can be avoided, and it is anticipated that failures will occur
under extreme conditions that exceed the code-required and utility-
established design criteria. A systematic investigation can provide informa-
tion that may be used to reaffirm or improve design criteria and
maintenance practices. The investigation may reveal that the conditions
were in excess of design criteria and no modification of the criteria or main-
tenance practices is justified. The goal of the failure investigation is to
establish the cause of the failure and try to reconstruct or understand the
behavior of the line subsequent to the failure initiation.
There has been much public reporting of failures in recent years, but
little has been published dealing with the technical aspects of transmission
line failures. Information on structural failure investigations may be found
in publications by Carper (1986) and Janney (1979), and in 1973 a series of
papers on transmission line failures (Griffing and Leavengood 1973) was
published.
The correct interpretation of the causes of transmission line failures has,
at times, led to significant modifications of line design practices. The inves-
tigator should be certain that the assumed failure mechanism is consistent
with the evidence.

195
196 Electrical Transmission Line Structural Loading

J.2 SAFETY

The site of a transmission line failure can be a deceptively dangerous


place, particularly before the restoration and repair crews have secured the
area. Although the broken structure may look stable enough, potential dan-
ger is all around with the possibility of wires still under tension and/or
energized, and structural members balanced in precarious positions.
Figures J-1 and J-2 are two examples of transmission line failures where
the site had to be secured prior to initiating the investigation. The cause of
failure in Figure J-1 was an extreme wind event. The structure and wires
had to be secured before failure investigators could approach the area and
proceed with their evaluation. The cause of failure in Figure J-2 was a
microburst. This is an obvious example of a failed structure to be avoided
until restoration crews are able to secure the site.
At all times, be aware of and follow the applicable utility safety proce-
dures. Document the site from a safe distance until the site is deemed safe
to approach either by experience and judgment of the investigator or by
the restoration and repair crews.

Figure J-1. Failure of a 230 kV lattice structure.


Appendix J 197

Figure J-2. Failure of a 46kV wood pole.

J.3 LEGAL CONSIDERATIONS FOR FAILURE INVESTIGATIONS

Fortunately, most transmission line failures do not result in legal actions.


However, in the rare event that personal injury, fatalities, or extensive dam-
age to private property are the result of a transmission line failure, the
failure investigation team must be aware that more rigorous processes and
procedures may be prudent and in some cases required as the data gath-
ered during the failure investigation could potentially be used in legal
proceedings.
The investigative team should be aware of and consider the following if it
appears that legal action may result from the failure(s) being investigated:
• Defer to all investigative activities of law enforcement, fire, medical,
safety, or other governmental agencies. The failure investigation
198 Electrical Transmission Line Structural Loading

being conducted on behalf of the utility should not proceed until


the site is released.
• Contact legal counsel to advise of the field situation and obtain
guidance in fact and data gathering and for contact with parties
outside of the utility.
• Take special care when gathering physical evidence. Document its
location with photographs prior to moving. Document handling
and storage of the physical evidence such that the “chain of cus-
tody” is maintained and evidence thoroughly documented.
• Engage a third-party engineering firm or laboratory to perform any
testing or analysis deemed necessary.
• In all instances, perform thorough and comprehensive data collec-
tion, perform state-of-the-art analysis, rigorously support any
conclusions, and behave professionally.

J.4 NEED FOR AND BENEFIT OF THOROUGH INVESTIGATIONS

In any failure event, the utility’s responsibility is to ensure public safety


and to promptly restore service. Therefore, a predefined emergency
response plan should be established so repair and restoration crews can be
mobilized quickly and a qualified engineer has adequate time to perform
a thorough investigation. Time is the vital factor and, unless plans have
been made before the event and priority directives issued, significant evi-
dence and data could be lost.
A utility/transmission line owner should have an established phone list
that identifies key failure investigation personnel. These individuals should
be familiar with the utility’s investigation procedures and policy. The list
should be distributed to the utility’s line construction and maintenance
office(s).
The reasons for attempting to get to the root causes of a failure event are
many:
• The cause may be an actual overload of ice, wind, or a combination
of the two that exceeded the design specifications and will require
an assessment of future risks and costs. The accurate assessment of
the actual ice and wind loads is imperative to determine whether
there was excessive loading or whether there was a problem or
defect within the system.
• Detection of a deficiency or defect may permit modifications to
components to prevent further failures or may lead to modifica-
tions of current design practices or specifications.
Appendix J 199

• The cause may be attributed to the deterioration of specific line


components that may justify increased inspection and replacement
policies.
• Unanticipated dynamic behavior may be detected.
• The investigation may uncover a specific loading case that was not
originally considered.
• A systematic and thorough failure investigation should provide the
line engineer a greater familiarity with the ways in which the
various components of the wire and structural support systems
interact when the system is severely stressed.

J.5 CAUSES OF FAILURE

The following lists represent some of the more general causes of trans-
mission line failures.

J.5.1 Natural Phenomena Exceeding Design Criteria


• Extreme wind,
• Extreme ice,
• Combination of ice and wind,
• Landslides,
• Avalanches,
• Ice movement on rivers or lakes (for structures located in the
water),
• Flooding (causing damage to structure or to foundation), and
• Soil liquefaction.

J.5.2 Human Causes


• Sabotage, vandalism, or theft of members and/or bolts; and
• Accidental damage caused by equipment and vehicles.

J.5.3 Structure Deficiencies (When Design Criteria Were Not Exceeded)


• Design inadequacies of structure and/or foundation,
• Missing members or loose bolts caused by vibration or omitted
during erection,
• Erroneously fabricated members,
• Improperly installed foundations, and
• Deterioration or corrosion of structures.
200 Electrical Transmission Line Structural Loading

J.5.4 Conductor, Ground-Wire, and Hardware Deficiencies


• Improper wire splices,
• Faulty or inadequate hardware,
• Fatigue failure of wire or hardware components,
• Insulation failures, and
• Deterioration or corrosion.

J.5.5 Construction-Related Causes


• Excessive vertical load during stringing,
• Excessive longitudinal load during stringing,
• Improper stringing sequence,
• Damage during stringing, and
• Inadequate inspection (of bolted connections, member alignment,
pole jacking, method of damping unstrung arms).

J.5.6 Improper Installation of Structures


• Manufacturer’s recommended slip joint jacking force was not
achieved,
• Field-assembled joints were not inspected prior to stringing,
• Unstrung arms were not dampened or tied off after installation,
• Structure sections were not properly aligned, and
• Manufacturer’s bolt tightening procedures were not followed.

J.6 FAILURE INVESTIGATIONS

A failure investigation can be a very simple and quick observation of the


facts represented by the evidence. At other times, it will result in a study
involving many engineers over a period of years. The least demanding of
investigations are those that follow an accident caused by an obvious
event, such as aircraft contact, foundation washout, and so forth. The
emphasis in the investigation will be directed to finding a means of pre-
venting recurrence and determining whether the post failure behavior of
adjacent structures was satisfactory.
A more difficult problem will be encountered when the cause can be
identified as a wind or ice storm but the evidence indicates that the failure
occurred at lower than the expected design values. These situations require
an examination of the evidence to determine whether there was a structure
design deficiency. For example, bolts or members may have been missing,
foundations may have had inadequate cover, or guy anchors may have had
inadequate uplift capacity. In other cases, consideration of yawed or lon-
Appendix J 201

gitudinal wind loads may have been omitted from the design criteria, or
probable uplift loads were not considered.
In the case of line damage with multiple failures caused by ice or wind
load equal to or exceeding design values, the investigation should attempt
to determine the line section that failed by the initial ice and/or wind
event. This should be inspected separately from other sections that may
have failed due to secondary events. This is an important finding to better
understand the behavior of the line, but it is often difficult to distinguish
between them.

J.7 PREPARATION

J.7.1 Failure Investigation Equipment


Following is a list of some of the essential items that the investiga­tive team
will want to have with them in the field as they conduct their inspection:
• Measuring tape and pocket scale,
• Micrometer (if material sizing is in question),
• Notebook and sketch pad,
• Markers and identification tags,
• Cardboard pieces or 8 × 11 in. paper pad and marker to place in
foreground of all photos for future identification,
• Voice recorder,
• High-resolution digital camera with video capability and extra batteries,
• Binoculars,
• Cell phone or radio, and
• Large sealable plastic bags to collect accumulated ice on wires.

J.7.2 Technical Preparation


If time and access permit, the investigators should familiarize them-
selves with the appropriate line data, conductor and structure loadings,
design characteristics, and any special construction records prior to con-
ducting the field investigation. When possible, they should discuss the
failure briefly with a group of key design personnel.

J.8 STEPS FOR AN EFFECTIVE FAILURE INVESTIGATION


CHECKLIST

The investigators should initiate complete photo documentation, make


an overall survey of the damaged area, and listen to viewpoints and evi-
dence of any witnesses or earlier arrivals.
202 Electrical Transmission Line Structural Loading

For utilities, first responders typically consist of patrolmen/linemen or


switching personnel. In addition to their primary responsibilities during a
transmission line failure, first responders are in an opportune position to
quickly collect initial data (site photographs are a good example) themselves
or from eyewitnesses that may have been on the scene at the time of the
failure(s). These reviews are invaluable as the incident is still fresh in every-
one’s mind. This also requires utilities to train their first responders on such
data gathering as well as provide them with time and encouragement to do so.
The following information may be used as a summary checklist for fail-
ure investigation but does not cover all the tasks that could be performed
during an investigation.

J.8.1 Data Gathering


Field data gathering items could include the following:
• Prior to arriving on site, confirm that the line is locked out (Lock
out-Tag out) of service at the substation and de-energized.
• Assume all wires or conductors are live unless confirmed by the
local utility. Downed conductors can still have lethal currents.
Utility-established practices for work around potentially energized
facilities must be followed.
• Develop a plan for conducting the investigation of the failure(s) in
the field. Evaluate the nature and extent of the failure(s) and based
on the size and available resources of the investigative team(s)
develop a plan to deploy to locations that appear to offer the best
opportunity to collect the highest quality and most informative
data. Communicate with the appropriate operations, maintenance,
or construction personnel/organizations to understand the status
of the failed facility and restoration plans. Seek out the first
responders to the failure for initial impressions and logistics
guidance. Provision the investigation team and initiate the field
investigation as quickly as practicable.
• The line restoration and repair crews may have already arrived at
the site and will be ready to start repair operations. If this happens,
try to obtain a visual inspection of the damaged portion of the line.
An overall picture taken at this time may provide information and
detail that could be lost after the repair activity begins. Observa-
tions of marks left by ground impacts may be useful in evaluating
the order of failures when secondary failures exist.
• The first impression of the site can result in a multitude of ideas
about the failure, and it is valuable for the investigator to record
these thoughts.
• Prepare a sketch of the line showing the positions of conductors,
insulators, structures, and any indication of the conductors having
Appendix J 203

been pulled across the ground. Also note the structure configura-
tion, such as the position of the guy anchors, deflected shape of
structures, and final position of footing stubs and/or structure legs.
• If the event is an ice storm, attempt to gather representative ice
samples from the fallen wires, record the length of each sample
along with the diameter of wire that it comes from, and store them
in plastic bags for later weighing. Sample ice weights are the best
way to accurately measure the ice load that was on the wires.
• An awareness of conductor and shield wire behavior is important
because these tie the structures together. Observe and document the
position of the wires/insulators in relation to the position of the
failed structure.
• If wind is the suspected cause of failure, look for surrounding
damage to trees, buildings, and so forth. The Beaufort scale (Bau-
meister et al. 1978), given in Table J-1, can provide valuable infor-
mation as to the approximate wind speed.
• Look for signs of the following:
ƒ Rust on sheared surfaces that may indicate that the bolt or
member had loss of cross section prior to the event,
ƒ Burn marks on the conductor or structure indicating initial point
of fault to ground,
ƒ Evidence of loose or missing bolts, and
ƒ Shiny steel and worn galvanizing at joints, indicating possible
vibration.
• If hardware, insulators, conductors, or overhead ground wires are
broken, they may have been triggered by the initial failure or may
have been caused by a secondary event. Retrieve and mark some
specimens as needed.
• If there are broken wires, note whether the ends of the strands
indicate a prior fracture due to fatigue, or a cup cone failure with
necking indicative of a tensile failure.
• It may be desirable to remove test sections of steel members for
material tests to determine material properties. Record the location
of the member samples. Avoid taking samples in the area of high
stress because the cold working of the steel will significantly alter
its physical properties. If a torch is used to remove the sample, be
sure to obtain a sample large enough that a testing coupon can be
prepared that has not been degraded due to the localized effects of
heat.
• Individuals in the nearby area of the failure may be a possible
source of information. These individuals can frequently tell of
vibration, galloping, and other unusual meteorological events that
may have occurred immediately prior to the failure or historically.
• Data should be maintained in a professional manner. Any data gath-
ered in the field, including photographs, may be discoverable in court.
204 Electrical Transmission Line Structural Loading

Table J-1.  Beaufort Scale of Wind Intensity

Wind Terms used in


Beaufort speed US Weather
number (mph) Wind effects observed on land Bureau reports
0 <1 Calm, smoke rises vertically. Light
1 1–3 Direction of wind shown by smoke
drift but not by wind vanes.
2 4–7 Wind felt on face, leaves rustle,
ordinary vane moved by wind.
3 8–12 Leaves and small twigs in constant Gentle
motion, wind extends light flag.
4 13–18 Raises dust, loose paper; small Moderate
branches are moved.
5 19–24 Small trees in leaf begin to sway, Fresh
crested wavelets form on inland
waters.
6 25–31 Large branches in motion, whis- Strong
tling heard in telegraph wires,
umbrellas used with difficulty,
wind is heard in buildings.
7 32–38 Whole trees in motion, difficult
walking against wind.
8 39–46 Branches break off trees, wind
generally impedes progress.
9 47–54 Slight structural damage occurs; Gale
chimney pots, slates removed.
10 55–63 Seldom experienced inland; trees Whole gale
uprooted, considerable structural
damage occurs, telephone poles
break.
11 64–72 Very rarely experienced, accompa-
nied by widespread damage.
12 > 73 Very rarely experienced, disastrous Hurricane
damage.
Source: Baumeister et al. (1978), US Weather Bureau.
Appendix J 205

Figure J-3 is an example of field data collection and the need to contact
multiple entities. The cause of the structure failure was a microburst. As
part of the failure investigation, first responders from utilities as well as
others such as railway officials should be approached for information.

Figure J-3. Failure of a 46 kV wood pole line.

J.8.2 After Returning to the Office


• Obtain weather data from the nearest local weather station.
• Look for additional weather recording stations within the local area
of the failure. Examples include substations, telecommunication or
cellular stations, highway bridges and overpasses, and homes or
businesses.
• Gather design criteria, “as-built” drawings, construction records,
and maintenance history of the line.
• When applicable, evaluate failed hardware or send these out for
expert evaluation.
• Weigh ice samples or measure the volume of melted ice and
calculate the ice load on the wire.
206 Electrical Transmission Line Structural Loading

J.8.3 Analysis of Data Gathered


• Look for design inadequacies:
ƒ Conductor weaker than structure,
ƒ Combined loading producing critical member stresses not
previously considered,
and
ƒ Foundation or anchor failures.
• Study field data carefully; try to match field data with postulated
cause of failure.
• If the structure appears to have failed below the design load, a
more detailed analysis may be warranted, taking into account
actual material yield strengths obtained by testing and secondary
stresses due to bending and nonlinearities.
• Contact the local utility operations personnel and determine the
exact time and nature of the line outage and any other relevant
information.
• Examine conductor behavior after the failure event and its potential
effect on the remaining transmission line system.
• Ascertain why damage terminated where it did.

J.8.4 Preparation of Report


The report should summarize and document the following:
• Field investigation, observations made, and data collected. The data
collected in the field (primarily photographs, sketches, interviews,
and notes) should be cataloged for future reference.
• Overview of the physical characteristics and layout of the line,
design practices, inspection methods, maintenance practices, and
construction techniques prior to the failure.
• Documentation of the failure summarizing the environmental
conditions, cause of the failure, identification of initial failure
location, sequence of failure, and contributing or mitigating factors.
• Conclusions and recommendations, including adequacy of design
criteria, inspection and maintenance practices, effectiveness of
failure containment, and recommendations for improvement or
modification of new or existing facilities.
• Follow-up evaluation of failure investigation.
Appendix J 207

J.9 POST-FAILURE BEHAVIOR OR FAILURE CONTAINMENT

The investigation should establish the cause of failure and whether the
line performed as designed. If needed, make recommendations regarding:
• Strengthening of the existing structures,
• Improvement of maintenance and inspection procedures,
• Possible change of load adjustments to design criteria and design
practices for future lines, and
• Failure containment.
Another function is to identify any evidence of a cascading failure. An
initial failure with collapsed structures or broken wires may cause damage
to one or two structures adjacent on either side. It is difficult to prevent
such damage in all cases because the nature of the initial event and the
impact and energy release may not be easily absorbed. If subsequent struc-
tures fail, a cascade is more likely. The investigator should also determine
the effectiveness of any existing failure containment measures.

J.10 ADDITIONAL REFERENCES: FAILURE INVESTIGATION


SPECIFIC

ASCE. 1989. “Guidelines for failure investigation.” New York: ASCE Task
Committee on Guideline for Failure Investigation, Technical Council on
Forensic Engineering.
ASCE. 1997. “Forensic engineering.” In Proc., 1st Congress, Forensic Engi-
neering Division of ASCE, Minneapolis, Minnesota. New York: ASCE.
ASCE. 2003. Guidelines for forensic engineering practice. Reston, VA: ASCE.
EPRI (Electric Power Research Institute). 2003. The fundamentals of forensic
investigation procedures guidebook, 1001890. Palo Alto, CA: EPRI.
EPRI. 2004. “Forensic analysis of failures.” Chapter 5 in Overhead transmis-
sion inspection and assessment guidelines—2004, 1002007. Palo Alto, CA:
EPRI.
EPRI. 2012. “Forensic analysis of failures.” Chapter 5 in Overhead transmis-
sion inspection and assessment guidelines—2012, 10024111. Palo Alto, CA:
EPRI.
APPENDIX K
HIGH-INTENSITY WINDS

K.1 INTRODUCTION

Load cases simulating the critical downburst and tornado wind configura-
tions for a generic transmission line system are provided in this appendix.

K.2 DOWNBURSTS

K.2.1 Proposed Critical Downburst Loads


As described in Section 2.2.1, the distribution and magnitude of forces
on the line components are a function of the downburst jet velocity VJ, the
practical jet diameter DJ, and the location of the downburst wind relative
to the target structure, as shown in Figure 2-12. A value of 110 to 160 mph
(50 to 70 m/s) is recommended for VJ.
Based on the results of extensive parametric studies conducted using an
experimentally validated numerical model, three load cases, providing an
envelope for the maximum effect of the downburst wind field on transmis-
sion line structures, were suggested by El Damatty and Elawady (2015).
One of the load cases requires carrying out nonlinear structural analysis
for the wire system. In lieu of nonlinear structural analysis, a set of charts
is provided in this appendix to estimate the conductor and shield wire
unbalanced longitudinal forces associated with this load case. A list of the
nomenclature is provided in Section K.4.

209
210 Electrical Transmission Line Structural Loading

Load Case 1: Transverse Wind Load (Ψ = 0°)  This case corresponds to a


downburst outflow in the transverse direction. The velocity profile
associated with this load case can be described as follows:
• Vertical distribution of the radial velocity along the height of the
structure, normalized with respect to the jet velocity VJ, is provided
in Figure K-1. An equivalent uniform velocity profile Veqt that
results in a base shear and overturning moment equal to or exceed-
ing those obtained from the downburst profile is calculated. It is
found that Veqt = 1.1VJ.
• Non-uniform distribution of the radial velocity along three spans
from each side of the structure of interest is shown in Figure K-2.
This symmetric distribution leads to a transverse force acting on the
structure of interest. It is found that an equivalent uniform distribu-
tion Veqc with a magnitude of 1.06VJ can be used to calculate the
transverse force on the wires.
Since the 3-second gust velocity is used as the reference velocity for down-
bursts, a span reduction factor can be applied to simulate the lack of correla-
tion in turbulence that is expected to occur along the conductor’s spans. The
relation between the conductor span and the reduction factor for the case of
downburst was developed by Holmes et al. (2008) and Behncke and Ho (2009).
This is compared to the span reduction factor for synoptic wind in Figure K-3.

Figure K-1. Radial velocity distribution over structure height for Ψ = 0°.
Source: El Damatty and Elawady (2018).
Appendix K 211

Figure K-2. Transverse radial velocity distribution over six conductor spans for Ψ = 0°.
Source: El Damatty and Elawady (2018).

Figure K-3. Downburst span reduction factor for peak pressures as a function of
average span length.
Source: Adapted from Holmes et al. (2008) and Behncke and Ho (2009).

Load Case 2: Longitudinal Wind Load (Ψ = 90°)  This case corresponds


to a downburst outflow in the longitudinal direction. Note that any wires
spanning between structures are not loaded in this case. The velocity
profile associated with this load case can be described as follows:
• The Vertical distribution of the radial velocity along the height of
the structure, normalized with respect to the jet velocity VJ, is
shown in Figure K-4. An equivalent uniform velocity profile Veqt
212 Electrical Transmission Line Structural Loading

that results in a base shear and overturning moment equal to or


exceeding those obtained from the downburst profile is calculated.
It is found that Veqt = 1.1VJ.
• No force acts on the conductors due to this load case.

Figure K-4. Longitudinal radial velocity distribution over structure height for
Ψ = 90°.
Source: El Damatty and Elawady (2018).

Load Case 3: Oblique Wind Load (Ψ = 30°)


This case is associated with a radial velocity profile that acts at an
oblique angle relative to the transmission line. The wind field associated
with this configuration will lead to the following effects:
• A velocity profile along the height of the structure in the transverse
direction (Y-direction), which can be approximated by an equiva-
lent uniform velocity VeqtT of 0.80VJ as shown in Figure K-5.
• A velocity profile along the height of the structure in the longitudi-
nal direction (X-direction), which can be approximated by an
equivalent uniform velocity VeqtL of 0.47VJ as shown in Figure K-5.
• Velocity profile that acts on the conductors, which has an unequal
distribution on the spans adjacent to the structure of interest as shown
in Figure K-6. This velocity profile leads to a transverse force, RY, as
well as a longitudinal unbalanced force, RX, acting on the structure.
Appendix K 213

Figure K-5. Radial velocity distribution over structure height for Ψ = 30°.
Source: El Damatty and Elawady (2018).

Figure K-6. Radial velocity distribution over six conductor spans for Ψ=30°.
Source: El Damatty and Elawady (2018).
214 Electrical Transmission Line Structural Loading

An equivalent uniform distribution, Veqc, with a magnitude of approxi-


mately 0.65VJ can be used to calculate the transverse force on the wires, RY,
under this load configuration. The evaluation of the longitudinal unbal-
anced force, RX, requires a non-linear analysis for the line wires. In order
to simplify the analysis, Elawady and El Damatty (2015) developed a set
of graphs that are provided in Section K.2.2. These graphs can be used to
evaluate the longitudinal unbalanced force of the conductors and shield
wires following the procedure described in the next section.
Because the downburst wind speeds are gust wind speeds, the gust
response factor, G, and velocity pressure exposure coefficient, Kz, should
be considered to be 1.0 for calculating the wind force.

K.2.2 Evaluation of Longitudinal Unbalanced Load RX for Downburst Case 3


K.2.2.1 Conductor Longitudinal Force  Due to the uneven distribution of
the loads acting on the conductor spans in this load case, the wire tension
forces on the spans adjacent to the target structure are different. This
difference is transferred to the structure in the form of unbalanced forces
acting in the longitudinal direction of the line. The evaluation of this
unbalanced longitudinal load requires performing nonlinear analysis for
the conductors, while taking the in-line swing of the insulators into account.
The main parameters that affect the magnitude of this longitudinal force
are the conductor’s projected diameter and weight, span length, conductor
sag, the length of suspension insulator, and the downburst equivalent jet
velocity (El Damatty et al. 2013).
To simplify these calculations, a set of charts is provided in this appen-
dix from which the longitudinal unbalanced force can be evaluated. Table
K-1 shows the upper and lower range of properties considered in the
developed charts. In this table, the square of the jet velocity VJ2 and the wire

Table K-1.  Parameters Considered in the Conductor Design Charts

Lower range Upper range


US customary US customary
Parameter SI units units SI units units
Insulator 1m 39.37 in. 5m 196.85 in.
length h
α 40 m3/sec2 0.1244 mile3/hr2 200 m3/sec2 0.6218 mile3/hr2
Wire weight w 10 N/m 0.6852 lb/ft 40 N/m 2.7408 lb/ft
Wire span L 100 m 328 ft 500 m 1640 ft
Sag ratio (sag/ 1m 39.37 in. 5m 196.85 in.
span ratio) S
Insulator 2% 2% 2% 2%
length h
Appendix K 215

projected diameter d were combined into one parameter, α. This parameter,


α, is directly proportional to the transverse loads acting on the conductors.
The magnitude of the longitudinal unbalanced force is found to vary lin-
early with the insulator length h while it varies nonlinearly with the other
parameters. However, it is found that by dividing the range of α into four
regions and the wire weight w into two regions, the magnitude of the lon-
gitudinal forces varies linearly with α, w, and h within each region. As
such, for each region, the charts are provided showing the variations of the
longitudinal force with the span and sag ratio for the extreme values of the
three parameters α, h, and w.
For each of the eight groups (I to VIII), eight graphs are provided for the
combinations of the maximum and minimum of the parameter values α,
h, and w. Each graph provides the variation of the longitudinal unbalanced
force with the span and the sag ratio (Elawady and El Damatty 2015). Lin-
ear interpolation can be done between the eight curves to obtain the lon-
gitudinal unbalanced force of the considered system as explained in the
following steps.
1. Calculate α = VJ2 × d.
2. Based on the values of α and w, determine to which group the
system belongs (I to VIII). Table K-2 provides simple guidance for
the selection of the design groups.

Table K-2.  Longitudinal Load Chart Guidance

α (m3/sec2) w (N/m)
Figure
αmin αmax wmin wmax Group reference
40 80 10 25 I K-7
40 80 25 40 II K-8
80 120 10 25 III K-9
80 120 25 40 IV K-10
120 160 10 25 V K-11
120 160 25 40 VI K-12
160 200 10 25 VII K-13
160 200 25 40 VIII K-14

3. Based on the span value and the sag ratio, determine the longitudi-
nal unbalanced force associated with the eight graphs of each
group. Those are labeled:
RX1 = Longitudinal force corresponding to (wmin, hmin, αmin)
RX2 = Longitudinal force corresponding to (wmin, hmin, αmin)
216 Electrical Transmission Line Structural Loading

RX3 = Longitudinal force corresponding to (wmin, hmin, αmin)


RX4 = Longitudinal force corresponding to (wmin, hmin, αmin)
RX5 = Longitudinal force corresponding to (wmin, hmin, αmin)
RX6 = Longitudinal force corresponding to (wmin, hmin, αmin)
RX7 = Longitudinal force corresponding to (wmin, hmin, αmin)
RX8 = Longitudinal force corresponding to (wmin, hmin, αmin)
Note: hmax = 5 m, wmax and αmax are the upper range for each group,
while hmin = 1 m, wmin and αmin are the lower range for each
group.
4. Based on the values of α, h, and w and the above eight evaluated
longitudinal forces, linear interpolation can be conducted using this
set of equations:
(a − amin )
RX ( 3−4 ) = RX 3 + (RX 4 − RX 3 )× (K-1)
(amax − amin )
(   min )
RX ( 5  6 )  RX 5  ( RX 6  RX 5 )  (K-2)
( max   min )
(a − amin )
RX (7 −8 ) = RX 7 + (RX 8 − RX 7 )× (K-3)
(amax − amin )

(wmax − w)
RX ( hmax) = RX (7 −8 ) + (RX ( 5−6 ) − RX (7 −8 ) )× (K-4)
(wmax − wmin )

(RX ( hmin ) − RX ( hmax ) )


RX = RX ( hmax ) + ×( hmax − h) (K-5)
( hmax − hmin )
( wmax  w)
RX ( hmax)  RX ( 7 8)  ( RX (56 )  RX ( 7 8) )  (K-6)
( wmax  wmin )

(a − amin ) (K-7)
RX ( 3−4 ) = RX 3 + (RX 4 − RX 3 )×
(amax − amin )

where Rx is the critical longitudinal unbalanced force resulting from


this downburst load case and is to be applied at the insulator to structure
connection.
5. The preceding analysis has been performed for a single conductor.
For cases of multiple conductor bundles, the longitudinal unbal-
anced force obtained from Equation (K-7) must be multiplied by the
number of subconductors in the bundle.
Appendix K
217

Figure K-7. Longitudinal force (RX) charts: Group I


218
Electrical Transmission Line Structural Loading

Figure K-7 (Continued). Longitudinal force (RX) charts: Group I.


Appendix K

Figure K-8. Longitudinal force (RX) charts: Group II.


219
220
Electrical Transmission Line Structural Loading

Figure K-8 (Continued). Longitudinal force (RX) charts: Group II.


Appendix K

Figure K-9. Longitudinal force (RX) charts: Group III.


221
222
Electrical Transmission Line Structural Loading

Figure K-9 (Continued). Longitudinal force (RX) charts: Group III.


Appendix K

Figure K-10. Longitudinal force (RX) charts: Group IV.


223
224
Electrical Transmission Line Structural Loading

Figure K-10 (Continued). Longitudinal force (RX) charts: Group IV.


Appendix K

Figure K-11. Longitudinal force (RX) charts: Group V.


225
226
Electrical Transmission Line Structural Loading

Figure K-11 (Continued). Longitudinal force (RX) charts: Group V.


Appendix K

Figure K-12. Longitudinal force (RX) charts: Group VI.


227
228
Electrical Transmission Line Structural Loading

Figure K-12 (Continued). Longitudinal force (RX) charts: Group VI.


Appendix K

Figure K-13. Longitudinal force (RX) charts: Group VII.


229
230
Electrical Transmission Line Structural Loading

Figure K-13 (Continued). Longitudinal force (RX) charts: Group VII.


Appendix K

Figure K-14. Longitudinal force (RX) charts: Group VIII.


231
232
Electrical Transmission Line Structural Loading

Figure K-14 (Continued). Longitudinal force (RX) charts: Group VIII.


Appendix K 233

K.2.2.2 Shield Wire Longitudinal Force  The main difference between a


shield wire and a conductor is that a shield wire is attached to the structure
directly or with a very short assembly. Separate design charts for the
longitudinal unbalanced forces are developed for the shield wires (Elawady
and El Damatty 2015). The value of the longitudinal forces are found to
depend on the parameters α, w, L, and S described previously, in addition
to the axial stiffness of the shield wire EA, where E is the modulus of
elasticity and A is the cross-sectional area of the wire. The range of shield
wire parameters considered are found in Table K-3.
Table K-3.  Parameters Considered in the Shield Wire Design Charts

Lower range Upper range


US customary US customary
Parameter SI units units SI units units
a 30 m3/sec2 0.0933 mile3/hr2 100 m3/sec2 0.311 mile3/hr2
Cable 3 N/m 0.205 lb/ft 15 N/m 1.028 lb/ft
weight w
Axial 0.9 × 107 N 2.02 × 106 lb 6 × 107 N 1.35 × 107 lb
stiffness EA
Cable span L 100 m 328 ft 500 m 1640 ft
Sag ratio 2.5% 2.5% 4% 4%
(sag/span
ratio) S

The following steps can be conducted to obtain the longitudinal force


associated with Load Case 3 for shield wires:
1. Calculate a = VJ2 × d.
2. Based on the span value and the sag ratio, determine the longitudinal
forces associated with the eight graphs of the design group (Fig-
ure K-15). Those are labeled
RX1 = Longitudinal force corresponding to (wmin, EAmin, αmin)
RX2 = Longitudinal force corresponding to (wmin, EAmin, αmin)
RX3 = Longitudinal force corresponding to (wmin, EAmin, αmin)
RX4 = Longitudinal force corresponding to (wmin, EAmin, αmin)
RX5 = Longitudinal force corresponding to (wmin, EAmin, αmin)
RX6 = Longitudinal force corresponding to (wmin, EAmin, αmin)
RX7 = Longitudinal force corresponding to (wmin, EAmin, αmin)
RX8 = Longitudinal force corresponding to (wmin, EAmin, αmin)
Note: EAmax = 6 × 107 N, EAmin = 0.9 × 107 N, wmax = 15 N/m,
wmin = 3 N/m, αmax = 100 m3/sec2, and αmin = 30 m3/sec2.
234
Electrical Transmission Line Structural Loading

Figure K-15. Longitudinal force (RX) charts: Shield wire group.


Appendix K

Figure K-15 (Continued). Longitudinal force (RX) charts: Shield wire group.
235
236 Electrical Transmission Line Structural Loading

3. Based on the shield wire values of α, EA, and w, and the eight
evaluated longitudinal forces, linear interpolation can be performed
using this set of equations:
(a − amin )
RX ( 5−6 ) = RX 5 + (RX 6 − RX 5 )× (K-8)
(amax − amin )

(a − amin )
RX (7 −8 ) = RX 7 + (RX 8 − RX 7 )× (K-9)
(amax − amin )

(wmax − w)
RX ( EAmax ) = RX (7 −8 ) + (RX ( 5−6 ) − RX (7 −8 ) )× (K-10)
(wmax − wmin )

(RX ( EAmax ) − RX ( EAmin ) )


RX = RX ( EAmin ) + ×(EA − EAmin ) (K-11)
(EAmax − EAmin )
(   min )
R X ( 7 8 )  R X 7  ( R X 8  R X 7 )  (K-12)
( max   min )
( wmax  w)
RX ( EAmax )  RX ( 7 8)  ( RX (56 )  RX ( 7 8) )  (K-13)
( wmax  wmin )
( RX ( EAmax )  RX ( EAmin ) )
RX  RX ( EAmin )   ( EA  EAmin ) (K-14)
( EAmax  EAmin )
where Rx is the critical longitudinal unbalanced force resulting from
this downburst load case and is to be applied at the shield wire to
structure connection.

K.2.3 Example: Downburst Load Case 3


This example shows calculations of the conductor longitudinal unbal-
anced force under oblique downburst loading. The loads are based on the
structure shown in Figure 5-1, the data are provided in Tables 5-1 to 5-3,
and the design data are as follows:

Design Data
• Wind span = 1,500 ft = 457.2 m;
• Length of insulator assembly = 6 ft = 1.83 m;
• Conductor self-weight = 1.075 lb/ft = 15.7 N/m;
• Conductor projected diameter = 1.165 in. = 0.03 m;
• Line sag = 36 ft = 11 m (~2.5% span); and
• Assumed downburst jet velocity of 112 mph = 50 m/s.
Appendix K 237

Based on the aforementioned data, the following calculations are


performed:

a = VJ2 × d = 502 × 0.03 = 75 m3/s2(0.2332 mi2/h2)

⇒ (40(0.1244 mi3/h2)) < a < (80(0.2487 mi3/hr2))

and w = 15.7 N/m (1.08 lb/ft)

⇒ (10(0.685 lb/ft)) < w < (25(1.713 lb/ft))

⇒ Group I, Figure K-7.

Based on the charts given in Figure K-7 for Group I, the following values
are extracted:
RX1 = 1.8 kips RX5 = 0.4 kips
RX2 = 7.3 kips RX6 = 1.0 kips
RX3 = 0.85 kips RX7 = 0.3 kips
RX4 = 3.0 kips RX8 = 0.78 kips

The following calculations are then performed:

(   min )
RX (1 2 )  RX 1  ( RX 2  RX 1 ) 

( max   min )
(0.2332 − 0.1244)
⇒ RX(1−2 ) = 1.8 + (7.2 − 1.8)× = 6.52 kips
(0.2487 − 0.1244)

(a − amin )
RX ( 3−4 ) = RX 3 + (RX 4 − RX 3 )×
(amax − amin )
(0.2332 − 0.1244)
⇒ RX( 3−4 ) = 0.85 + (3.0 − 0.85)× = 2.68 kips
(0.2487 − 0.1244)
( wmax  w)
RX ( h min)  RX (3 4 )  ( RX (1 2 )  RX (3 4 ) ) 
( wmax  wmin )
(1.713 − 1.08)
⇒ RX ( h min) = 2.68 + (6.52 − 2.68)× = 5.061 kips
(1.713 − 0.685)
(a − amin )
RX ( 5−6 ) = RX 5 + (RX 6 − RX 5 )×
(amax − amin )

238 Electrical Transmission Line Structural Loading

(0.2332 − 0.1244)
⇒ RX( 5−6 ) = 0.4 + (1.0 − 0.4)× = 0.92 kips
(0.2487 − 0.1244)

(a − amin )
RX (7 −8 ) = RX 7 + (RX 8 − RX 7 )×
(amax − amin )
(0.2332 − 0.1244)
⇒ RX(7−8 ) = 0.3 + (0.78 − 0.3)× = 0.73 kips
(0.2487 − 0.1244)
(wmax − w)
RX ( h max) = RX (7 −8 ) + (RX ( 5−6 ) − RX (7 −8 ) )×
(wmax − wmin )
(1.713 − 1.08)
⇒ RX ( h max) = 0.73 + (0.92 − 0.73)× = 0.84 kiips
(1.713 − 0.685)

Finally, the longitudinal unbalanced force is evaluated as follows:
(RX ( h min) − RX ( h max) )
RX = RX ( h max) + ×( hmax − h)
( hmax − hmin )

(5.061  0.84)
 RX  0.84   (16.4  6)  4.18 kips
(16.4  3.28)

K.3 TORNADOES

K.3.1 F2 Tornado Critical Load Cases


Two methods of analysis to account for the critical effect of tornadoes
on transmission line structures are provided in this appendix. It should be
noted that the simplified analysis is applicable only for cantilever-type
structures (self-supported), while the detailed analysis can be applied to
any transmission line system.

K.3.1.1 Simplified Analysis  For cantilever-type structures (self-supported),


the critical effect of tornadoes is simulated by applying a uniform velocity
along the height of the structure with a magnitude of 161 mph (72 m/sec)
accompanied with a uniform velocity of 161 mph (72 m/sec) applied on
the wires. This uniform velocity field is applied from all potentially
sensitive directions. The effect of the yaw angle Ψ should follow the
procedure described in Section 2.1.6.2. The forces acting on the wires due
to this uniform velocity can be reduced using the span reduction factor
(SRF) calculated using the Equation (K-15) (Behncke and Ho 2009):
Appendix K 239

SRF = WG (1 − 0.25WG/L)/L (K-15)

where WG is the tornado gust width with a recommended value of


490 ft (150 m) for F2 tornadoes, and L is the average span length of the
transmission line.

K.3.1.2 Detailed Analysis  Two simplified load cases that simulate the
critical effect of F2 tornadoes on transmission structures, which can be
applied to both self-supported and guyed structures, are provided as
follows (El Damatty et al. 2015). Each load case has two vertical velocity
profiles along the structure height combined with a uniformly distributed
transverse wind velocity profile on the conductors as shown.

K.3.1.2.1 Load Case 1  This load case is based on the following three wind
profiles:
1. Vertical velocity profile A acting on one face of the structure shown
in Figure K-16.
2. Vertical velocity profile B acting on the perpendicular face of the
structure shown in Figure K-17.
3. Profile C of uniform velocity distribution on the conductors with a
gust value of 161 mph (72 m/sec). The forces acting on the wires
due to this uniform velocity can be reduced using the span reduc-
tion factor (SRF) calculated using Equation (K-15).
It is not required to consider all potential wind directions in this load
case. Only the wind loading profile combinations shown in Figure K-18 are
to be considered in the application of this load case.

K.3.1.2.2 Load Case 2  This case corresponds to a tornado center located on


an oblique angle relative to the line. This load case is based on the follow-
ing three wind profiles:
1. Vertical velocity profile D acting on one face of the structure shown
in Figure K-19.
2. Vertical velocity profile E acting on the perpendicular face of the
structure shown in Figure K-20.
3. Profile F of uniform velocity distribution on the conductors with an
amplitude 161 mph (72 m/sec). The forces acting on the wires due
to this uniform velocity can be reduced using the span reduction
factor (SRF) calculated using Equation (K-15).
It is not required to consider all potential wind directions in this load
case. Only the wind loading profile combinations shown in Figure K-21 are
to be considered in the application of this load case.
240 Electrical Transmission Line Structural Loading

Figure K-16. Vertical Velocity Profile A.

Figure K-17. Vertical Velocity Profile B.


Appendix K 241

Figure K-18. Possible combinations of the vertical wind profiles A and B for
Load Case 1.

Figure K-19. Vertical Velocity Profile D.


242 Electrical Transmission Line Structural Loading

Figure K-20. Vertical Velocity Profile E.

Figure K-21. Possible combinations of the vertical wind profiles D and E.


Appendix K 243

For a structure, the following combinations are to be considered:


Tornado wind speeds are gust wind speeds; therefore, the gust response
factor, G, and velocity pressure exposure coefficient, Kz, should be consid-
ered equal to 1.0 for calculating the wind force.
Transmission structures may be susceptible to wind-borne debris initi-
ated by HIW. These load cases do not account for debris impact loads.

K.4 HIW NOMENCLATURE

K.4.1 Downburst Section


d Single wire projected diameter perpendicular to the radial
direction of the downburst velocity;
DJ Downburst jet diameter;
h Insulator length;
R Distance from the center of the downburst to the center of the
structure of interest;
RX Downburst critical longitudinal unbalanced force based on the
proposed design charts corresponding to the actual insulator
length;
RX (hmax) Wire longitudinal force corresponding to the selected design
group maximum insulator length, the line’s actual α, and the
line’s actual w;
RX (hmin) Wire longitudinal force corresponding to the selected design
group minimum insulator length, the line’s actual α, and the
line’s actual w;
RX1 Wire longitudinal force under Downburst Critical Case 3 corre-
sponding to specific design group (wmin, hmin, αmin);
RX2 Wire longitudinal force under Downburst Critical Case 3 corre-
sponding to specific design group (wmax, hmax, αmax);
RX3 Wire longitudinal force under Downburst Critical Case 3 corre-
sponding to specific design group (wmin, hmin, αmin);
RX4 Wire longitudinal force under Downburst Critical Case 3 corre-
sponding to specific design group (wmax, hmax, αmax);
RX5 Wire longitudinal force under Downburst Critical Case 3 corre-
sponding to specific design group (wmin, hmin, αmin);
RX6 Wire longitudinal force under Downburst Critical Case 3 corre-
sponding to specific design group (wmax, hmax, αmax);
RX7 Wire longitudinal force under Downburst Critical Case 3 corre-
sponding to specific design group (wmin, hmin, αmin);
RX8 Wire longitudinal force under Downburst Critical Case 3 corre-
sponding to specific design group (wmax, hmax, αmax);
244 Electrical Transmission Line Structural Loading

RX(1-2) Wire longitudinal force corresponding to the selected design


group minimum weight, minimum insulator length, and the
line’s actual α;
RX(3-4) Wire longitudinal force corresponding to the selected design
group maximum weight, minimum insulator length, and the
line’s actual α;
RX(5-6) Wire longitudinal force corresponding to the selected design
group minimum weight, maximum insulator length, and the
line’s actual α;
RX(7-8) Wire longitudinal force corresponding to the selected design
group maximum weight, maximum insulator length, and the
line’s actual α;
RY Wire force in the transverse direction;
S Wire sag divided by line span;
Veqc Equivalent uniform velocity distribution for wires;
Veqt Equivalent uniform velocity distribution for structures;
VJ Downburst jet velocity;
VRD Radial component of the downburst outflow velocity;
VVR Vertical component of the downburst outflow velocity;
w Single wire weight;
X-axis Axis that passes along the line direction (longitudinal direction);
Y-axis Axis perpendicular to the line direction (transverse direction);
α Product of VJ2 and d; and
Ψ Angle between the vertical plane of the transverse direction and
the vertical plane connecting the center of the downburst and the
center of the structure of interest.

K.4.2 Tornado Section


L Average span length of the transmission line;
r Radial distance from tornado center; and
WG Tornado gust width.
APPENDIX L
WEATHER-RELATED LOADS FOR
ADDITIONAL MRIS

This Manual of Practice recommends that transmission lines be designed


for weather-related loads (i.e., wind and ice) corresponding to a basic or
reference mean recurrence interval (MRI) of 100 years. It is acknowledged
that it may be of interest to design a transmission line to a level of reliability
lower or higher than the reference MRI. This appendix provides the design
data for selected MRIs and equations for loading to be consistent with
Chapter 2.
Maps of basic wind speeds were developed for 50-year and 300-year
MRIs (ASCE 2017) based on a methodology consistent with that discussed
in Chapter 2 (Pintar et al. 2015). Similar maps of radial ice thickness from
freezing rain with concurrent gust wind speeds were developed for 50-year
and 300-year MRI, also based on a methodology consistent with that dis-
cussed in Chapter 2. It should be pointed out that all the notes, methods of
application, and limitations apply to these maps as to those presented in
Chapter 2, and users should familiarize themselves with these aspects.

L.1 50-YEAR MRI

The basic wind speed map for a 50-year MRI is shown in Figure L-1.
Values are 3-second gust wind speeds in miles per hour (m/s also shown)
at 33 ft (10 m) above ground in terrain with Exposure Category C. This map
of wind speeds is associated with an annual exceedance probability of 0.02
(2%) and corresponds to approximately a 64% probability of exceedance in
50 years.
The entire state of Hawaii is defined as a Special Wind Region on the
current wind speed maps. This is due to the extreme topographic conditions

245
246 Electrical Transmission Line Structural Loading

found throughout the Hawaiian Islands. Following a review of the Kzt


maps for Hawaii, the following wind speeds are recommended for a MRI
of 50 years:
1. Wind speed of 82 mph (37 m/s) for regions indicated as Kzt ≤ 1.5
( ) ( )
2. Wind speed of 67 mph • K zt or 30 m / s • K zt for regions
indicated as Kzt > 1.5.
In areas where local historical icing data are not available, the glaze ice
maps given in Figures L-2 through L-6 can be used with some limitations.
These maps show 50-year MRI ice thicknesses due to freezing precipitation
with concurrent 3-second gust wind speeds in miles per hour (m/s) at 33 ft
(10 m) above ground for the continental United States and Alaska. The
glaze ice thicknesses shown in these figures do not include in-cloud icing
or sticky snow accretions, which are caused by meteorological conditions
that may produce significantly different loading patterns (see Appendix H,
Section H.5).

L.2 300-YEAR MRI

The basic wind speed map for a 300-year MRI is shown in Figure L-7.
Values are 3-second gust wind speeds in miles per hour (m/s) at 33 ft
(10 m) above ground in terrain with Exposure Category C. This map of
wind speeds is associated with an annual exceedance probability of 0.00333
(0.333%) and corresponds to approximately a 15% probability of exceed-
ance in 50 years.
The entire state of Hawaii is defined as a Special Wind Region on the
current wind speed maps. This is due to the extreme topographic condi-
tions found throughout the Hawaiian Islands. Following a review of the
Kzt maps for Hawaii, the following wind speeds are recommended for a
MRI of 300 years:
1. Wind speed of 141 mph (63 m/s) for regions indicated as Kzt ≤ 1.5
( ) ( )
2. Wind speed of 115mph • K zt or 51 m / s • K zt for regions
indicated as Kzt > 1.5.
In areas where local historical icing data are not available, the glaze ice
maps given in Figures L-8 through L-12 can be used with some limitations.
These maps show 300-year ice thicknesses due to freezing precipitation
with concurrent 3-second gust wind speeds in miles per hour (m/s) at 33
ft (10 m) above ground for the continental United States and Alaska. The
glaze ice thicknesses shown in these figures do not include in-cloud icing
or sticky snow accretions, which are caused by meteorological conditions
that may produce significantly different loading patterns (see Appendix H,
Section H.5).
Appendix L 247

Figure L-1. 50-year MRI 3-second gust wind speed map [mph (m/s)] at 33 ft
(10 m) aboveground in Exposure Category C.
Source: Minimum Design Loads and Associated Criteria for Buildings
and Other Structures (ASCE (2017).
248 Electrical Transmission Line Structural Loading

(a)

Figure L-2. 50-year MRI radial ice thickness (in.) from freezing rain with
concurrent gust wind speeds (mph) at 33 ft (10 m) aboveground in Exposure C:
(a) western United States; and (b) eastern United States. (Continued)
Appendix L 249

(b)

Figure L-2. (Continued) 50-year MRI radial ice thickness (in.) from freezing
rain with concurrent gust wind speeds (mph) at 33 ft (10 m) aboveground in
Exposure C: (a) western United States; and (b) eastern United States.
250 Electrical Transmission Line Structural Loading

Figure L-3. 50-year MRI radial ice thickness (in.) from freezing rain with
concurrent gust wind speeds (mph) at 33 ft (10 m) aboveground in Exposure C:
Puget Sound detail.

Figure L-4. 50-year MRI radial ice thickness (in.) from freezing rain with
concurrent gust wind speeds (mph) at 33 ft (10 m) aboveground in Exposure C:
Columbia River Gorge detail.
Appendix L 251

Figure L-5. 50-year MRI radial ice thickness (in.) from freezing rain with
concurrent gust wind speeds (mph) at 33 ft (10 m) aboveground in Exposure C:
Lake Superior detail.

Figure L-6. 50-year MRI radial ice thickness (in.) from freezing rain with
concurrent gust wind speeds (mph) at 33 ft (10 m) aboveground in Exposure C:
Alaska.
252 Electrical Transmission Line Structural Loading

Figure L-7. 300-year MRI 3-second gust wind speed map [mph (m/s)] at 33 ft
(10 m) aboveground in Exposure Category C.
Source: ASCE (2017).
Appendix L 253

(a)

Figure L-8. 300-year MRI radial ice thickness (in.) from freezing rain with
concurrent gust wind speeds (mph) at 33 ft (10 m) aboveground in Exposure C:
(a) western United States; and (b) eastern United States. (Continued)
254 Electrical Transmission Line Structural Loading

(b)

Figure L-8. (Continued) 300-year MRI radial ice thickness (in.) from freezing
rain with concurrent gust wind speeds (mph) at 33 ft (10 m) aboveground in
Exposure C: (a) western United States; and (b) eastern United States.
Appendix L 255

Figure L-9. 300-year MRI radial ice thickness (in.) from freezing rain with
concurrent gust wind speeds (mph) at 33 ft (10 m) aboveground in Exposure C:
Puget Sound detail.

Figure L-10. 300-year MRI radial ice thickness (in.) from freezing rain with
concurrent gust wind speeds (mph) at 33 ft (10 m) aboveground in Exposure C:
Columbia River Gorge detail.
256 Electrical Transmission Line Structural Loading

Figure L-11. 300-year MRI radial ice thickness (in.) from freezing rain with
concurrent gust wind speeds (mph) at 33 ft (10 m) aboveground in Exposure C:
Lake Superior detail.

Figure L-12. 300-year MRI radial ice thickness (in.) from freezing rain with
concurrent gust wind speeds (mph) at 33 ft (10 m) aboveground in Exposure C:
Alaska.
APPENDIX M
DRAFT PRE-STANDARD MINIMUM DESIGN
LOADS FOR ELECTRICAL TRANSMISSION
LINE FACILITIES

M.1 Purpose
M.2 Scope
M.3 Applicable Documents
M.4 Definitions
M.5 Notations
M.6 Load Cases for Strength Design
M.6.1 Basic Load Cases
M.6.2 Supplemental and Serviceability Load Cases
M.6.3 Load Factors
M.6.4 Reliability Adjustment
M.6.5 Strength
M.7 Dead Loads
M.8 Wire Loads
M.8.1 General
M.8.2 Dynamic Wire Loads
M.8.3 Unbalanced Longitudinal Loads
M.9 Wind Loads
M.9.1 General
M.9.2 Wind Force
M.9.2.1 Air Density Coefficient, Q
M.9.2.2 Basic Wind Speed
M.9.2.2.1 Special Wind Regions
M.9.2.2.2 Estimation of Basic Wind Speeds from
Regional Climatic Data
M.9.2.3 Limitations
M.9.2.4 Exposure Categories
M.9.2.5 Wind Pressure Exposure Coefficient

257
258 Electrical Transmission Line Structural Loading

M.9.2.6 Gust Response Factor


M.9.2.7 Force Coefficient
M.9.2.7.1 Wires
M.9.2.7.2 Lattices Truss Structures
M.9.2.7.3 Pole Strucures
M.9.2.8 Topographic Effects
M.10 Ice with Concurrent Wind Loads
M.10.1 General
M.10.1.1 Site-Specific Studies
M.10.2 Nominal Ice Thickness
M.10.3 Loads Due to Freezing Rain With Concurrent Wind
M.10.3.1 Design Ice Thickness for Freezing Rain
M.10.3.2 Ice Weight on Wires
M.10.3.3 Ice Weight on Structures and Non-Structural
Attachments
M.10.3.4 Wind on Structures and Non-Structural
Attachments
M.10.3.5 Wind on Ice-Covered Wires
M.10.3.6 Design Temperatures for Freezing Rain
M.10.4 Unbalanced Ice Loading
M.11 Legislated Loads
M.12 Construction and Maintenance Loads
M.12.1 General
M.12.2 Climbing Loads
M.13. Other High Consequence Events

M.1 PURPOSE

The ASCE Task Committee on Structural Loadings envisions the need


for a loading standard for transmission line facilities. This appendix pres-
ents the recommendations of this committee written in a prescriptive form
and presented as a Draft Pre-Standard for public review and comment.
Comments should be directed to the ASCE Committee on Electrical Trans-
mission Structures.

M.2 SCOPE

This Draft Pre-Standard provides minimum load requirements for the


design of transmission line facilities. Appropriate loads, load factors, and
loading combinations that have been developed to be used concurrently
are set forth for ultimate strength design (LRFD) using the referenced
design standards and guides.
Appendix M 259

This Draft Pre-Standard shall be used by the Transmission Owner or


their authorized agent responsible for developing the structural load cases
for the design of transmission line facilities. The minimum voltage that
applies to a transmission line shall be established by the Transmission
Owner.
NOTE: This Draft Pre-Standard is intended for design of new transmis-
sion line facilities. It may be applied to the assessment of existing facilities.
Further, these principles may be applied to temporary or emergency facili-
ties with adjustments.
NOTE: These principles may be applied to distribution facilities.

M.3 APPLICABLE DOCUMENTS

The following standards, codes, and guidelines are referenced in this


Draft Pre-Standard; the latest revisions apply unless noted:
• ACI 318, Building Code Requirements for Structural Concrete and
Commentary;
• AISC 360, Specification for Steel Buildings;
• ANSI C2, National Electrical Safety Code (NESC);
• ANSI C29, Series of standards for both ceramic and non-ceramic
insulators;
• ANSI C119, Series of standards for conductors;
• ANSI O5, Series of standards for wood products;
• ASCE 7-16, Minimum Design Loads and Associated Criteria for Build-
ings and Other Structures;
• ASCE 10, Design of Latticed Steel Transmission Structures;
• ASCE 48, Design of Steel Transmission Pole Structures;
• ASCE MOP 74, Guidelines for Electrical Transmission Line Structural
Loading;
• ASCE MOP 91, Design of Guyed Electrical Transmission Structures;
• ASCE MOP 104, Recommended Practice for Fiber-Reinforced Polymer
Products for Overhead Utility Line Structures;
• ASCE MOP 123, Prestressed Concrete Transmission Pole Structures;
• ASCE MOP 141, Wood Pole Structures for Electrical Transmission
Lines;
• ASCE Guide for the Design and Use of Concrete Poles;
• IEEE Standard 524, IEEE Guidelines to the Installation of Overhead
Transmission Line Conductors;
• IEEE Standard 951, IEEE Guide to the Assembly and Erection of Metal
Transmission Structures;
• IEEE Standard 1025, IEEE Guide to the Assembly and Erection of
Concrete Pole Structures;
260 Electrical Transmission Line Structural Loading

• IEEE 1307, IEEE Standard for Fall Protection for Electric Utility Trans-
mission and Distribution on Poles and Structures; and
• OSHA Occupational Safety and Health Administration applicable
regulations.

M.4 DEFINITIONS

Basic Wind Speed: The 3-second gust wind speed at 33 ft (10 m) above
ground in open country terrain (Exposure Category C).
Effective Height: The theoretical height above ground to the center of pres-
sure of the wind load.
Exposure Category: A description of the terrain features and ground
roughness upwind of the transmission line facility.
Force Coefficient: A coefficient accounting for the effects of member char-
acteristics (e.g., shape, size, solidity, shielding, orientation with respect
to the wind, surface roughness) on the resultant force due to wind. It is
also referred to as the drag coefficient, pressure coefficient, or shape
factor.
Freezing Rain: Rain or drizzle that falls into a layer of subfreezing air at
the earth’s surface and freezes on contact with the ground or an object
to form glaze ice.
Glaze Ice: Clear high-density ice, with a density of approximately 56 pcf
(900 kg/m3).
Gust Response Factor: The ratio of the peak load effect on the structure or
wires to the mean load effect corresponding to the design wind speed.
It is a multiplier of the design wind load to obtain the peak load effect.
Hoarfrost: An accumulation of ice crystals formed by direct deposition of
water vapor from the air onto an object.
Ice Accretion: The formation of ice on transmission line facilities and non-
structural attachments.
In-Cloud Icing: Ice occurring when supercooled cloud or fog droplets car-
ried by the wind freeze on impact with objects. In-cloud icing usually
forms rime, but may also form glaze.
Longitudinal: Local axis of the structure that is generally parallel to the
direction of the wires and perpendicular to the vertical axis.
Mean Recurrence Interval (MRI): The inverse of the probability of exceed-
ance of an environmental load (i.e., wind, ice) in any given year. For
example, a design event with the probability of exceedance of 0.01 (1%)
in any given year is associated with an MRI of 100 years.
Nonstructural Attachments: Components attached to the structure or
wires with mass or surface area that significantly contributes to the over-
all structural loading. Such attachments include, but are not limited to,
Appendix M 261

electrical and communication equipment, signs, ladders, platforms,


aerial marker balls, bird diverters, and anti-galloping devices.
Rime Ice: White or opaque ice with entrapped air. Typical densities range
from 20 to 50 pcf (320 to 800 kg/m3).
Snow: Snow that adheres to objects by some combination of capillary
forces, freezing, and sintering. The snow may be either wet or dry. Typi-
cal densities range from 20 to 60 pcf (320 to 960 kg/m3).
Solidity Ratio: A ratio of the area of all members in the windward face of
a latticed structure (net area) to the area of the outline of the windward
face of a latticed structure (gross area).
Transmission Line Facilities: All wires, insulators, hardware, supporting
structures, guy wires, anchors, and foundations.
Transverse: Local axis of the structure that is generally perpendicular to
the direction of the wires and the vertical axis.
Vertical: Local axis of the structure that is parallel with the direction of the
gravitational force.
Wind Pressure Exposure Coefficient: A coefficient used to modify the
basic wind pressure to account for variations of wind speed with height
due to interaction (friction) with the surface roughness of the earth.
Wire: All electrical conductors, shield wires, optical ground wire, messen-
gers, and communication cables attached to a transmission structure.
Yawed Wind: Wind at angles of incidence to the transmission structure or
line other than the longitudinal and transverse loading directions.

M.5 SYMBOLS

The following symbols are used in this Draft Pre-Standard:


A = Area projected on a plane normal to the wind direction [ft2 (m2)];
Am = Area of all members in the windward face of a latticed structure
(net area) [ft2 (m2)];
Ao = Area of the outline of the windward face of a latticed structure
(gross area) [ft2 (m2)];
Bt = Dimensionless response term corresponding to the quasi-static
background wind loading on the structure;
Bw = Dimensionless response term corresponding to the quasi-static
background wind loading on the wires;
cexp = Turbulence intensity constant, based on exposure;
Cf = Force coefficient associated with the windward face of the
structure;
d = Diameter of bare wire [inches (mm)];
F = Wind force in the direction of wind unless otherwise specified
[lb (N)];
G = Gust response factor for structures and wires;
262 Electrical Transmission Line Structural Loading

Gt = Gust response factor for the structure;


Gw = Gust response factor for the wires;
HIW = High Intensity Wind
Iz = Turbulence intensity at the effective height of the tower/struc-
ture or wire;
Kz = Wind pressure exposure coefficient, which modifies the reference
wind pressure for various heights above ground based on
different exposure categories;
Kzt = Topographic factor;
LC = Load Case
Ls = Integral length scale of turbulence [ft (m)];
MRI = Mean Recurrence Interval
Q = Air density coefficient in the wind force equation that converts the
kinetic energy of moving air into potential energy of pressure;
QD = Design load effect in each component of a structure;
Qi = Constant to convert radial ice thickness to weight;
QMRI = Reliability adjustment factor;
S = Design wind span [ft (m)] of the wires;
t = Nominal ice thickness due to freezing rain at a height of 33 ft
(10 m) [inches (mm)];
tMRI = Nominal ice thickness due to freezing rain at a height of 33 ft
(10 m) at a selected MRI [inches (mm)];
tz = Design ice thickness [inches (mm)];
t100 = Nominal ice thickness attributable to freezing rain at a height of
33 ft (10 m) for 100-year MRI from Figures M-2 through M-7
[inches (mm)];
VMRI = Reference 3-second gust wind speed for selected MRI [mph (m/s)];
V100 = Reference 3-second gust wind speed for 100-year MRI from
Figure M-1 [mph (m/s)];
w = Wire weight per unit length [lb/ft (N/m)];
Wi = Weight of glaze ice per unit length [lb/ft (N/m)];
z = Height aboveground [ft (m)];
zg = Gradient height, which defines the thickness of the atmospheric
boundary layer [ft (m)]; above this height, the wind speed is
constant;
zh = Effective height from ground level to the center of pressure of the
wind load on the structure, or effective height of the wire [ft (m)];
α = Power law coefficient for gust wind;
γ = Load factor appropriate for the event;
Φ = Solidity ratio (Am/Ao);
ϕ·Rn = Design resistance or deflection restriction of the transmission line
facility; and
Ψ = Yaw angle measured in a horizontal plane, referenced as 0° for
wind perpendicular to the wires (degrees).
Appendix M 263

M.6 LOAD CASES FOR STRENGTH DESIGN

The Transmission Owner or authorized agent shall design transmission


line facilities with sufficient strength for the basic load cases defined in
Section M.6.1. Consideration shall also be given to the supplemental and
serviceability load cases defined in Section M.6.2.
Load cases shall be multiplied by the applicable load factors defined in
Section M.6.3.
Where transmission line facilities warrant a reliability level different from
that defined in this Draft Pre-Standard (i.e., MRI100) due to site-specific
application, the provisions of Section M.6.4 shall be followed.
The load cases of this section are cumulatively represented by the
following:

∑ (γ·LC·QMRI) (M-1)

where
LC = Load Case defined in Sections M.6.1 and M.6.2,
γ = Load factor appropriate for the event defined in Section
M.6.3, and
QMRI = Reliability adjustment factor defined in Section M.6.4.

M.6.1 Basic Load Cases


Transmission line facilities shall be designed such that their design
strength equals or exceeds the effects of the following load cases. Determi-
nation of the magnitude of each load, including the effects of dead and wire
loads, shall be in accordance with the applicable section of this Draft
Pre-Standard.
• Climatic Load Cases
ƒ Extreme wind load
ƒ Extreme ice load with concurrent wind load and temperature
• Line Security Load Cases
ƒ Unbalanced longitudinal load
ƒ Failure containment load
• Operational Load Cases
ƒ Construction load with concurrent weather condition
ƒ Maintenance load with concurrent weather condition
ƒ Equipment operation load when applicable
• Legislated Loads
264 Electrical Transmission Line Structural Loading

M.6.2 Supplemental and Serviceability Load Cases


Where site-specific circumstances warrant, the following load cases shall
be considered in the design of transmission line facilities. These loads may
not have an associated MRI:
• Load criteria associated with deflection limitations.
• Dynamic wire loading with associated weather event.
NOTE: Dynamic loading should be evaluated relative to appropri-
ate component resistance considering the nature of the loading.
• Earthquake events
NOTE: The structural capacity provided by designing transmission
line facilities to the loading requirements of this Draft Pre-Standard
provides sufficient capability to resist earthquake ground motions
(inertia loads). Site and soil conditions and foundation types should
be reviewed for projects in high seismic regions to address potential
failures from secondary events such as soil liquefaction and
landslides.
• Other potential high consequence events such as floods, tsunamis, snow
creep, and avalanches. See Section M.13 for additional information.

M.6.3 Load Factors


Unless otherwise specified within this Draft Pre-Standard, the following
load factors (γ) shall be the minimum used. Any unique situation deemed
by the Transmission Owner or authorized agent to warrant a load factor
greater than 1.0 shall be considered.
• 1.0: All weather-related loads.
Note: ASCE Manual of Practice No. 74 prefers designers use MRI to
adjust resistance to weather events based on the importance of
specific facility. See Section M.6.4.
• 1.5: Construction and maintenance loads.
• 1.0: Dead load, structure weight, and weight of supported facilities.
• Legislated loads: Use the load factors as defined by the legislated
document.

M.6.4 Reliability Adjustment


The MRI used in the calculation of climatic loads may be adjusted for
transmission line facilities requiring a reliability level different from that
defined within this Draft Pre-Standard (i.e., MRI100). This may be due to
site-specific applications or studies, or operating circumstances such as
those described below:
• Challenging site access and restoration circumstances
Examples: River, highway, railroad crossing;
Appendix M 265

• Temporary construction;
• Emergency restoration; and
• The importance of the line relative to the performance of the
transmission grid or service provided.
The Transmission Owner or authorized agent shall determine the appro-
priate MRI for those situations warranting one different from that specified
as the baseline of this Draft Pre-Standard. Wind speed and ice accretion
maps for other MRIs can be found in Appendix L of ASCE Manual of Prac-
tice No. 74 or ASCE 7-16.
Note: Climatic load cases as defined in this Draft Pre-Standard are deter-
mined by statistical modeling techniques. As such, the structural reliability
level can be adjusted.

M.6.5 Strength
The resistance of a transmission line facility shall exceed the effects of the
prescribed loads in this document as described by the following formula:

ϕ·Rn ≥ ∑(γ·LC·QMRI) (M-2)

where
ϕ·Rn = design resistance or deflection restriction of the transmission
line facility as defined by the appropriate design guide for
the applicable structure type or an appropriate serviceability
restriction.

The strength of transmission line facilities is beyond the scope of this


Draft Pre-Standard. The Transmission Owner or authorized agent is
referred to the following documents for the applicable material strength
factors and design requirements:
• ACI 318, Building Code Requirements for Structural Concrete and
Commentary;
• AISC 360, Specifications for Steel Buildings;
• ANSI C2, National Electrical Safety Code;
• ANSI C29, Series of standards for both ceramic and non-ceramic
insulators;
• ANSI C119, Series of standards for conductors;
• ANSI O5, Series of standards for wood;
• ASCE Standard No. 10, Design of Latticed Steel Transmission
Structures;
• ASCE Standard No. 48, Design of Steel Transmission Pole Structures;
• ASCE Manual of Practice No. 91, Design of Guyed Electrical Transmis-
sion Structures;
266 Electrical Transmission Line Structural Loading

• ASCE MOP 104, Recommended Practice for Fiber-Reinforced Polymer


Products for Overhead Utility Line Structures;
• ASCE MOP 123, Prestressed Concrete Transmission Pole Structures;
and
• ASCE MOP 141, Wood Pole Structures for Electrical Transmission Lines.
For strengths of other transmission line facilities such as insulators and
conductors, refer to the specifications of the applicable supplier.

M.7 DEAD LOADS

The weight of the structure, wires, and components such as insulators,


hardware, electrical equipment, and non-structural attachments shall be
included in the design of transmission line structures and foundations.

M.8 WIRE LOADS

M.8.1 General
The loads induced by all attached wires known at the time of initial
design shall be included in the design of the structure. Wire loads shall be
calculated based on tensions, span lengths, and line angles appropriate for
the site and for the temperature, ice, and wind loadings specified in this
Draft Pre-Standard. The effects of wind and ice on non-structural attach-
ments attached to wires shall be used in the calculation of design wire
tensions. Further, wire loads shall be applied in the various combinations
defined in Section M.6. Climatic, construction, and legislated loads shall
be included and combined as appropriate to determine the maximum load
effect.

M.8.2 Dynamic Wire Loads


Dynamic wire loads, such as those resulting from galloping, ice shed-
ding, and aeolian vibration that are caused or enhanced by wind and ice
or flexible structures and supports, shall be considered.

M.8.3 Unbalanced Longitudinal Loads


Unbalanced longitudinal loads resulting from unequal tensions on adja-
cent spans due to variation in icing or wind speed; broken shield wire,
conductor, or sub-conductors; conductor stringing; and secondary loading
resulting from a structural failure shall be considered.
Appendix M 267

M.9 WIND LOADS

M.9.1 General
Transmission line facilities shall be designed to resist the wind loads
determined in accordance with this section. Wind loading with all other
applicable loads addressed herein shall be applied in the direction that
produces the maximum load effect.

M.9.2 Wind Force


The wind force acting on the projected surface area of components of
transmission line facilities and non-structural attachments shall be deter-
mined by Equation (M-3a) or (M-3b)

F = QKzKzt(V100)2GCfA (M-3a)

or

F = QKzKzt(VMRI)2GCfA (M-3b)

where
F = Wind force in the direction of wind unless otherwise specified
[lb (N)];
G = Gust response factor for structures and wires as specified in
Section M.9.2.6;
Cf = Force coefficient as defined in Section M.9.2.7;
A = Area of the component projected on the plane normal to the
wind direction [ft2 (m2)];
Q = Air density coefficient defined in Section M.9.2.1;
Kz = Wind pressure exposure coefficient which modifies the
reference wind pressure for various heights above ground
based on different exposure categories; the values are
obtained from Section M.9.2.5;
Kzt = Topographic factor; 1.0 unless the guidance in Section M.9.2.8
and the procedures of ASCE 7 are followed;
V100 = Reference 3-second gust wind speed for 100-year MRI [mph
(m/s)] obtained from Figure M-1 in Section M.9.2.2; and
VMRI = Reference 3-second gust wind speed for selected MRI [mph
(m/s)] obtained from ASCE 7-16 or Appendix L of ASCE
Manual of Practice 74.
268 Electrical Transmission Line Structural Loading

M.9.2.1 Air Density Coefficient, Q  For wind speed in miles per hour
(m/s) and pressure in pounds per square foot (Pa), Q is defined in Equation
(M-4). A different value of Q may be used if justified by analysis of site-
specific elevation and temperature data. Refer to ASCE Manual of Practice
No. 74, Appendix C for additional information.

Q = 0.00256 customary units (0.613 metric units) (M-4)

M.9.2.2 Basic Wind Speed  The basic wind speed associated with a 100-year
MRI, V100, used in the determination of design wind loads on transmission
line facilities and non-structural attachments shall be determined from Figure
M-1, except as provided in Sections M.9.2.2.1 and M.9.2.2.2. Linear
interpolation between contours is permitted. The last wind speed contour of
the coastal area may be used for islands and coastal areas beyond the last
contour.
If wind speeds associated with a MRI other than 100 years are required,
the designer is referred to the additional wind speed maps in Appendix L
of ASCE Manual of Practice No. 74 or ASCE 7-16.

M.9.2.2.1 Special Wind Regions  Mountainous terrain, gorges, or other spe-


cial wind regions shown in Figure M-1 shall be examined for unusual wind
conditions. If necessary, the designer shall adjust the values given in Fig-
ure M-1 to account for local wind speeds. Such adjustment shall be based
on meteorological information and an estimate of the basic wind speed
obtained in accordance with the provisions of Section M.9.2.2.2.
For transmission line facilities in the state of Hawaii, the basic wind
speed shall be determined as follows:

105 mph (47 m/s) for regions indicated as Kzt ≤ 1.5

(86 mph • K zt ) or (38 m / s • K zt ) for regions indicated as K


zt > 1.5

Values of Kzt shall be determined using the appropriate wind map avail-
able from the Department of Accounting and General Services for the state
of Hawaii.

M.9.2.2.2 Estimation of Basic Wind Speeds from Regional Climatic Data  In


areas outside hurricane-prone regions, regional climatic data shall only be
used in lieu of the basic wind speeds given in Figure M-1 when (1)
approved extreme value statistical analysis procedures have been employed
in reducing the data; and (2) the length of record, sampling error, averaging
time, anemometer height, data quality, and terrain exposure of the ane-
mometer have been taken into account. The extreme value statistical analy-
Appendix M 269

sis may be used to justify a reduction in basic wind speed below that of
Figure M-1.
The use of regional wind speed data obtained from anemometers is not
permitted to define the hurricane wind speed risk along the hurricane-prone
regions of the continental United States, Hawaii, Puerto Rico, Guam, the
Virgin Islands, and American Samoa. In hurricane-prone regions, wind
speeds derived from simulation techniques shall only be used in lieu of the
basic wind speeds given in Figure M-1 when
1. Wind industry accepted simulation procedures are applied (i.e.,
Monte Carlo simulations based on historical hurricane records).
2. An appropriate number of years of synthetic hurricane activity are
simulated and validated using historical key hurricane statistics.
NOTE: A minimum database of 50,000 to 100,000 years of simulated
hurricane activity is typically considered acceptable.
3. A wind engineering industry-accepted wind field model is used to
generate wind speeds based on hurricane track records.
4. Wind industry accepted extreme value statistical analysis proce-
dures are used for the estimation of extreme wind speeds (i.e., Type
I extreme value distribution) at a given location.
In areas outside hurricane-prone regions, when the basic wind speed is
estimated from regional climatic data, the basic wind speed shall not be
less than the wind speed associated with the specified mean recurrence
interval, and the estimate shall be adjusted for equivalence to a 3-second
gust wind speed at 33 ft (10 m) above ground in Exposure Category C.

M.9.2.3 Limitations  The load effects resulting from localized convective


winds, also referred to as high intensity winds (HIWs), such as thunderstorms,
downbursts, and tornadoes, are not addressed by this Draft Pre-Standard. Refer
to ASCE Manual of Practice No. 74 for additional information related to HIW.

M.9.2.4 Exposure Categories  Exposure Category C shall be used to design


transmission facilities, unless the criteria of Exposures B or D can be met and
the exposure category will not change over the life of the transmission line.
Exposure Category B is classified as urban and suburban areas, densely
wooded areas, or terrain with numerous, closely spaced obstructions hav-
ing the size of single-family dwellings or larger. Use of Exposure Category
B shall be limited to wind directions for which representative terrain
extends either 2,600 feet (792 m) or 20 times the height of the transmission
structure, whichever is greater.
In the use of Exposure Category B, the longest distance of flat, unob-
structed terrain located in the middle of a suburban area permitted before
Exposure Category C must be used is 600 ft (180 m) or 20 times the height
of the transmission structure, whichever is less.
270 Electrical Transmission Line Structural Loading
Appendix M 271

Figure M-1. 100-year MRI 3-second gust wind speed map [mph (m/s)] at 33 ft
(10 m) above ground in Exposure C.
Source: ASCE (2017).
272 Electrical Transmission Line Structural Loading

Exposure Category C is defined as open terrain with scattered obstruc-


tions having heights less than 30 ft (9.1 m). This category includes flat, open
country, farms, and grasslands. This exposure category should be used
whenever terrain does not fit the descriptions of the other exposure catego-
ries. This exposure category may be considered representative of airport
terrain, where most wind speed measurements are recorded.
Exposure Category D is described as flat, unobstructed areas directly
exposed to wind flowing over open water for a distance of at least 5,000 ft
(1,524 m), or 20 times the height of the structure, whichever is greater. This
exposure category applies to shorelines in hurricane-prone regions, inland
waterways, the Great Lakes, and coastal areas of California, Oregon, Wash-
ington, and Alaska. This exposure category applies only to structures
directly exposed to bodies of water and coastal beaches. Exposure Cate-
gory D extends inland from the shoreline a distance of 1500 ft (457 m) or
10 times the height of the transmission structure, whichever is greater.
For a site located in the transition zone between exposure categories, or
where the exposure category is determined to be different on opposite
sides of the transmission facility, the category resulting in the largest wind
forces shall be used.
Exposure Categories for the state of Hawaii shall be determined using
the appropriate wind map available from the Department of Accounting
and General Services for the state of Hawaii.
NOTE: For additional information related to the definition of Exposure
Categories and how to apply them to facilities in the transition zones
between categories, refer to Chapter 2 of ASCE Manual of Practice No. 74.

M.9.2.5 Wind Pressure Exposure Coefficient


The wind pressure exposure coefficient, Kz, used in Equation (M-1) shall
be calculated using Equation (M-5)

2
 z a

Kz = 2.01  h  for 33 ft ≤ zh ≤ zg (M-5)
 z g 

where
α = Power law coefficient for gust wind from Table M.9-1
zh = Effective height from ground level to the center of pressure of
the wind load, and
zg = Gradient height from Table M.9-1

For heights up to 200 ft (60 m) from ground level, Kz may be determined


using the values shown in Table M.9-2.
Appendix M 273

NOTE: For structural design purposes, the effective heights of all wires
may be approximated as the average height above ground of all the wire
attachment points to the structure. For structure heights 200 ft (60 m) or
less, the design engineer may assume the structure is comprised of one
section, and two-thirds of the total structure height may be used as the
effective height.

Table M.9.1.  Power Law Exponent for Gust Wind Speed and
Corresponding Gradient Height

Exposure α zg (ft)

B 7.0 1200
C 9.5 900
D 11.5 700

Table M.9-2.  Wind Pressure Exposure Coefficient, Kz

Wind pressure exposure coefficient, Kz


Effective height,
zh (ft) Exposure B Exposure C Exposure D
0–33 0.72 1.00 1.18
40 0.76 1.04 1.22
50 0.81 1.09 1.27
60 0.85 1.14 1.31
70 0.89 1.17 1.35
80 0.93 1.21 1.38
90 0.96 1.24 1.41
100 0.99 1.27 1.43
120 1.04 1.32 1.48
140 1.09 1.36 1.52
160 1.13 1.40 1.55
180 1.17 1.43 1.59
200 1.20 1.46 1.62
274 Electrical Transmission Line Structural Loading

M.9.2.6 Gust Response Factor


Structure and wire gust response factors, Gt and Gw, respectively, shall
be determined using Equations (M-6) and (M-7).
 1 + 4.6 I z Bt 
Gt =  (M-6)
 1 + 6.1I z 

 1 + 4.6 I z Bw 
Gw =  (M-7)
 1 + 6.1I z 

1
in which  33 6
I z = cexp   (M-8)
 z 
h

1
Bt = (M-9)
0.56 zh
1+
Ls

1
Bw = (M-10)
0.8S
1+
Ls

where
Bt = Dimensionless response term corresponding to the
quasi-static background wind loading on the structure,
Bw = Dimensionless response term corresponding to the
quasi-static background wind loading on the wires,
cexp = Turbulence intensity constant based on exposure and found
in Table M.9-3,
Iz = Turbulence intensity at effective height of the tower/structure
or wire,
Ls = Integral length scale of turbulence [ft (m)] found in Table M.9-3,
S = Design wind span [ft (m)] of the wires, and
zh = Two-thirds of the total height of the structure for the calcula-
tion of Gt using Equation (M-6), or effective height of the wire
for the calculation of Gw using Equation (M-7).
Table M.9-3.  Turbulence Parameters for Calculation
of Gust Response Factor by Exposure

Exposure cexp Ls (ft)


B 0.3 170
C 0.2 220
D 0.15 250
Appendix M 275

The complete gust response factor equations in Appendix F of ASCE


Manual of Practice No. 74 shall be considered for the calculation of Gt and
Gw when

Gt = Structure natural frequency is less than 1 Hz, (which typically


occurs with structures taller than 200 ft);
Gw = Horizontal span is longer than 2000 ft.

M.9.2.7 Force Coefficient  For members with aspect ratios greater than or
equal to 40, the force coefficient shall be determined from Sections M.9.2.7.1
through M.9.2.7.3. For members with aspect ratios less than 40, the
correction factors defined in Appendix G of ASCE Manual of Practice
No. 74 shall be applied.
For force coefficients of shapes not described within this Draft Pre-
Standard, refer to Appendix G of the ASCE Manual of Practice No. 74.

M.9.2.7.1 Wires  The force coefficient for single and bundled conductors
and for ground wires shall be as defined in Equation (M-11) unless more
definitive data based on wind force measurements are available.

Cf = 1.0 (M-11)

M.9.2.7.2 Latticed Truss Structures  Force coefficients for square-section and


triangular-section latticed truss structures having flat-sided members shall
be determined using Table M.9-4.
Table M.9-4.  Force Coefficients, Cf, for Normal Wind on Latticed Truss
Structures Having Flat-Sided Members

Tower cross section Cf


Square 4.0Φ2 − 5.9Φ + 4.0
Triangular 3.4Φ2 − 4.7Φ + 3.4
Source: Adapted from ASCE 7-16 (2017).

In Table M.9-4, Φ, the solidity ratio, is defined as


A
Φ = m (M-12)
Ao

where
Am = area of all members in the windward face of the structure (net
area) [ft2 (m2)], and
Ao = area of the outline of the windward face of the structure
(gross area) [ft2 (m2)].
276 Electrical Transmission Line Structural Loading

When the truss members consist of round members, a correction factor,


equal to 0.51Φ2 + 0.57, but not greater than 1.0, shall be multiplied by the
value determined in Table M.9-4.
The solidity ratio for selected panels in the transverse and longitudinal
faces shall be used for the determination of wind loads.
The force coefficient calculated above accounts for loads accumulated
by both the windward and leeward tower faces (including shielding
effects) and shall be applied directly in Equation (M-3).
When the wind is yawed, the force coefficient Cf shall be multiplied by
the wind angle magnification factor defined in Equation (M-13).

1 + 0.2sin2(2Ψ) (M-13)

where
Ψ = yaw angle, measured in the horizontal plane, and is refer-
enced as 0° for wind perpendicular to the wires.

Alternatively, when wind is applied to individual members of latticed


structures, force coefficients shall be 1.0 for components with round surfaces
and 1.6 for components with flat surfaces. Refer to Appendix G of ASCE
Manual of Practice No. 74 for values for member shapes not listed herein.
NOTE: Calculating the force coefficient using Table M.9-4 is used when
applying the wind load using the “Wind on Face” method. The alternate
method using the force coefficients of each individual truss member is used
when applying wind load using the “Wind on Member” method.
NOTE: For latticed structures consisting of multiple columns and beams
to form bent or bay structures, the force coefficients of each column or
beam shall be calculated independently without considering the shielding
effects of other members unless special wind studies show otherwise.

M.9.2.7.3 Pole Structures  Force coefficients for pole structures are given in
Table M.9-5. Refer to Appendix G of ASCE Manual of Practice No. 74 for
values for member shapes not listed.
Table M.9-5.  Force Coefficients, Cf, for Members of Pole Structures

Member shape Cf Adapted from


Circular 0.9 ASCE 7-16 (2017)
16-sided polygonal 0.9 James (1976)
12-sided polygonal 1.0 James (1976)
8-sided polygonal 1.4 ASCE 7-16 (2017), James (1976)
6-sided polygonal 1.4 ASCE 7-16 (2017)
Square, rectangle 2.0 ASCE 7-16 (2017)
Appendix M 277

M.9.2.8 Topographic Effects  In complex or mountainous terrain where


wind speeds vary dramatically with exposure or by channeling of wind,
special studies shall be used to determine the wind flow.
For wind speed-up due to flow over isolated hills and escarpments, the
procedures of ASCE Standard 7-16 shall be followed.

M.10 ICE WITH CONCURRENT WIND LOADS

M.10.1 General
Atmospheric ice loads due to freezing rain, snow, and in-cloud icing
shall be considered in the design of electric transmission line facilities.
Design ice thickness tz shall be no less than the nominal ice thickness result-
ing from a 100-year MRI.
In areas where records or experience indicate that snow or in-cloud icing
produces larger loads than from freezing rain, site-specific studies shall be
used. Structural loads due to hoarfrost are not a design consideration.

M.10.1.1 Site-Specific Studies  Site-specific studies shall be used to deter­


mine the 100-year MRI ice thickness, concurrent wind speed, and con­
current temperature in:
1. Alaska
2. Areas where records or experience indicate that snow or in-cloud
icing produces larger loads than freezing rain.
3. Special icing regions shown in Figures M-4, M-5, and M-6.
4. Areas where experience indicates unusual icing conditions exist.
In lieu of using the mapped values, it shall be permitted to determine
the ice thickness, the concurrent wind speed, and the concurrent tempera-
ture for transmission line facilities from local meteorological data based on
a 100-year MRI provided that:
1. The quality of the data for wind and type and amount of precipita-
tion has been taken into account.
2. A robust ice accretion algorithm, with accreted ice assumed to
remain on wires until the air temperature is above freezing, has
been used to estimate ice thicknesses and concurrent wind speeds
from these data.
3. Extreme-value statistical analysis procedures based on an extreme
value distribution with at least three parameters have been
employed in analyzing the ice thickness and concurrent wind speed
data.
4. The length of record and sampling error has been taken into
account.
278 Electrical Transmission Line Structural Loading

M.10.2 Nominal Ice Thickness


Nominal ice thickness shall be determined using Figures M-2 through
M-7. These figures show the equivalent uniform radial thicknesses of ice t
due to freezing rain at a height of 33 ft (10 m) over the contiguous 48 states
and Alaska for a 100-year MRI. Also shown are concurrent 3-second gust
wind speeds. Thicknesses for Hawaii, and for ice accretions due to other
sources in all regions, shall be obtained from local meteorological
studies.

M.10.3 Loads Attributable to Freezing Rain with Concurrent Wind

M.10.3.1 Design Ice Thickness for Freezing Rain  The design ice thickness
tz shall be calculated from Equations (M-14), (M-15), (M-16), or (M-17), as
appropriate.

tz = t100(z/33)0.10 for 0 ft < z < 900 ft (M-14)

or

tz = tMRI(z/33)0.10 for 0 ft < z < 900 ft (M-15)

In SI:

tz = t100(z/10)0.10 for 0 m < z < 275 m (M-16)

or

tz = tMRI(z/10)0.10 for 0 m < z < 275 m (M-17)

where:
t100 = Nominal ice thickness due to freezing rain at a height of 33 ft
(10 m) for 100-year MRI, from Figures M-2 through M-7 [in.
(mm)],
tMRI = Nominal ice thickness due to freezing rain at a height of 33 ft
(10 m) at a selected MRI [inches (mm)],
tz = Design ice thickness [in. (mm)], and
z = Height above ground [ft (m)]. For heights above ground
greater than 900 ft (275 m), use z = 900 ft (275 m).
Appendix M 279

M.10.3.2 Ice Weight on Wires


The ice weight, Wi, shall be determined using Equation (M-18) consider-
ing a uniform thickness of glaze ice on the full length of all wires.

Wi = Qi (d + tz) tz (M-18)

where:
Wi = Weight of glaze ice [lb/ft (N/m)],
Qi = Constant to convert ice thickness to weight, 1.24 in customary
units (0.0282 in metric units),
d = Diameter of bare wire [inches (mm)], and
tz = Design ice thickness [inches (mm)].

The ice density used in Qi shall not be less than 56 pcf (900 kg/m3).

M.10.3.3 Ice Weight on Structures and Non-Structural Attachments  Ice


weight on transmission structures and non-structural attachments need
not be considered for design of transmission structures. The weight of ice
accretion due to freezing rain on non-structural attachments on wires such
as aerial marker balls, bird diverters, and anti-galloping devices shall be
considered in the calculation of the design wire tensions.

M.10.3.4 Wind on Structures and Non-Structural Attachments  The wind


pressure shall be determined using the concurrent wind speeds in Figures
M-2 to M-7 and the procedures of Section M.9.1. The additional projected
area due to ice accretion may be neglected when calculating the wind force
on transmission structures and non-structural attachments.

M.10.3.5 Wind on Ice-Covered Wires  Wind pressures applied to ice-


covered wires shall be determined using the concurrent wind speeds in
Figures M-2 to M-7 and the procedures of Section M.9.1.

M.10.3.6 Design Temperatures for Freezing Rain The design


temperatures concurrent with the design ice and wind-on-ice loads due to
freezing rain shall be the temperature for the site shown in Figures M-8 and
M-9. Use either of these values or 32 °F (0 °C), whichever results in the
maximum load effect. A temperature of 32 °F (0 °C) shall be used in Hawaii.
These temperatures are applicable for ice thicknesses for all mean
recurrence intervals.
280 Electrical Transmission Line Structural Loading

Figure M-2. 100-year MRI radial ice thickness (in.) from freezing rain with
concurrent gust wind speeds (mph) at 33 ft (10 m) aboveground in Exposure C:
Western United States.
Appendix M 281

Figure M-3. 100-year MRI radial ice thickness (in.) from freezing rain with
concurrent gust wind speeds (mph) at 33 ft (10 m) aboveground in Exposure C:
Eastern United States.
282 Electrical Transmission Line Structural Loading

Figure M-4. 100-year MRI radial ice thickness (in.) from freezing rain with
concurrent gust wind speeds (mph) at 33 ft (10 m) aboveground in Exposure C:
Puget Sound detail.

Figure M-5. 100-year MRI radial ice thickness (in.) from freezing rain with
concurrent gust wind speeds (mph) at 33 ft (10 m) aboveground in Exposure C:
Columbia River Gorge detail.
Appendix M 283

Figure M-6. 100-year MRI radial ice thickness (in.) from freezing rain with
concurrent gust wind speeds (mph) at 33 ft (10 m) aboveground in Exposure C:
Lake Superior detail.

Figure M-7. 100-year MRI radial ice thickness (in.) from freezing rain with
concurrent gust wind speeds (mph) at 33 ft (10 m) aboveground in Exposure C:
Alaska.
284 Electrical Transmission Line Structural Loading

Figure M-8. Temperatures concurrent with ice thickness attributable to freezing


rain (contiguous 48 states).

Figure M-9. Temperatures concurrent with ice thickness attributable to freezing


rain: Alaska.
Appendix M 285

M.10.4 Unbalanced Ice Loading


The effects of unbalanced ice loading on the wires in adjacent spans shall
be considered.

M.11 LEGISLATED LOADS

Transmission line facilities shall be designed to resist the loading speci-


fied by all applicable federal, state, and municipal codes, and legislative or
administrative acts.

M.12 CONSTRUCTION AND MAINTENANCE LOADS

M.12.1 General
Loading from construction and maintenance activities shall be consid-
ered in the design of transmission line facilities. NOTE: For additional
information on installation procedures, refer to the latest revisions of the
following documents:
• IEEE Standard 524, IEEE Guide to the Installation of Overhead Trans-
mission Line Conductors;
• IEEE Standard 951, IEEE Guide to the Assembly and Erection of Metal
Transmission Structures;
• IEEE Standard 1025, IEEE Guide to the Assembly and Erection of
Concrete Pole Structures; and
• Transmission Owner’s construction specifications.

M.12.2 Climbing Loads


Loads acting on the structures resulting from workers installing or
maintaining transmission facilities shall be considered in the design of the
structure. Climbing anchorage attachment points and loads shall be con-
sidered for worker safety.
Minimum loading shall meet or exceed all Occupational Safety & Health
Administration (OSHA) and other governmental requirements as applicable,
and the latest revisions of IEEE 1307 the “IEEE Standards for Fall Protection for
Utility Work” IEEE Standard 1307, and ANSI C2, National Electrical Safety Code.

M.13. OTHER HIGH-CONSEQUENCE EVENTS

Events with the potential for devastating damage to transmission line


facilities resulting in extended outages shall be considered. Examples of these
events include tornadoes, floods, landslides, avalanches, and sabotage.
REFERENCES

AASHTO. 1975. Standard specifications for structural supports for highway


signs, luminaires, and traffic signals. Washington, DC: AASHTO.
Abbey, R. E., Jr. 1976. “Risk probabilities associated with tornado
wind-speeds.” In Proc., Symp. on Tornadoes: Assessment of Knowledge and
Implications for Man, Lubbock, Texas: Texas Tech Univ., 177–236.
Abild, J., E. Y. Andersen, and L. Rosbjerg. 1992. “The climate of extreme
winds at the Great Belt, Denmark.” J. Wind Eng. Ind. Aerodyn. 41 (1–3):
521–532. https://doi.org/10.1016/0167-6105(92)90458-M.
Aboshosha, H., and A. El Damatty. 2013. “Span reduction factor of trans-
mission line conductors under downburst winds.” In Proc., 8th
Asia-Pacific Conf. on Wind Engineering, Chennai, India. https://doi.
org/10.3850/978-981-07-8012-8_P11.
ACI (American Concrete Institute). 2014. Building code requirements for struc-
tural concrete and commentary. ACI 318-14. Farmington Hills, MI: ACI.
AISC. 2010. Specification for structural steel buildings. AISC 360-10. Chicago:
AISC.
ALCOA (Aluminum Company of America). 1961. Graphic method for sag
tension calculations for ACSR and other conductors. Pittsburgh: ALCOA.
Anjo, K., S. Yamasaki, Y. Matsubayashi, Y. Nakayama, A. Otwsuki, and T.
Fujimura. 1974. “An experimental study of bundle conductor galloping
on the Kasatori-Yama test line for bulk power transmission.” In Proc.,
CIGRÉ 25th Session, Vol. 1 (22–04), Paris.
ANSI (American National Standards Institute). 2002. Specifications and
dimensions for wood poles. ANSI O5.1. New York: ANSI.
Armitt, J., M. Cojan, C. Manuzio, and P. Nicolini. 1975. “Calculation of
wind loadings on components of overhead lines.” In Proc., Inst. Electr.
Eng. 122 (11): 1247–1252. https://doi.org/10.1049/piee.1975.0306.

287
288 References

Arya, S. P. S., M. E. Capuano, and L. C. Fagen. 1987. “Some fluid modeling


studies of flow and dispersion over two dimensional hills.” Atmos. Envi-
ron. 21 (4): 753–764. https://doi.org/10.1016/0004-6981(87)90071–0.
ASCE. 1961. “Wind forces on structures.” Trans. ASCE 126 (2): 1124–1198.
ASCE. 1982. “Loadings for electrical transmission structures.” J. Struct.
Div.108 (ST5): 1088–1105.
ASCE. 1990a. Design of steel transmission pole structures. ASCE Manual of
Practice No. 52. New York: ASCE.
ASCE. 1990b. Minimum design loads for buildings and other structures. ASCE
7-88. New York: ASCE.
ASCE. 1991. Guidelines for electrical transmission line structural loading. ASCE
Manual of Practice No. 74, 2nd ed. New York: ASCE.
ASCE. 2000. Design of latticed steel transmission structures. ASCE 10-15. Res-
ton, VA: ASCE.
ASCE. 2002. Dynamic response of lattice towers and guyed masts. Edited by M.
K. S. Madugula. Reston, VA: ASCE.
ASCE. 2005. Minimum design loads for buildings and other structures. ASCE
7-05. Reston, VA: ASCE.
ASCE. 2006. Reliability-based design of utility pole structures. ASCE Manual
of Practice No. 111. Reston, VA: ASCE.
ASCE. 2010a. Guidelines for electrical transmission line structural loading.
ASCE Manual of Practice No. 74, 3rd ed. Reston, VA: ASCE.
ASCE. 2010b. Minimum design loads for buildings and other structures. ASCE
7-10. Reston, VA: ASCE.
ASCE. 2011. Design of steel transmission pole structures. ASCE 48-11. Reston,
VA: ASCE.
ASCE. 2017. Minimum design loads and associated criteria for buildings and
other structures. ASCE 7-16. Reston, VA.
Baenziger, M. A., S. Gupta, T. J. Wipf, F. Fanous, Y. H. Hahm, and H. B.
White. 1993a. Structural failure analysis of 345 kV transmission line, and
discussion by B. White. SM 441-6 PWRD. New York: IEEE.
Baenziger, M. A., W. D. James, B. Wouters, L. Li, and H. B. White. 1993b.
Dynamic loads on transmission line structures due to galloping conductors,
and discussion by B. White. WM 078-6 PWRD. New York: IEEE.
Baumeister, T., E. A. Avallone, T. Baumeister III, eds. (1978). Standard hand-
book for mechanical engineers, 8th ed. New York: McGraw-Hill, 9−167.
Bayar, D. C. 1986. “Drag coefficients of latticed towers.” J. Struct. Eng. 112 (2):
417–430. https://doi.org/10.1061/(ASCE)0733-9445(1986)112:2(417).
BEAIRA (British Electrical and Allied Industries Research Association).
1935. “Wind pressure on latticed towers—Tests on models.” J. Inst.
Electr. Eng. 77 (464): 189–196. https://doi.org/10.1049/jiee-1.1935.0139.
Behncke, R. H., and T. C. E. Ho. 2009. “Review of span and gust factors for
transmission line design.” In Proc., Electrical Transmission and Substation
Structures 2009, Fort Worth, TX. https://doi.org/10.1061/41077(363)18.
References 289

Bennett, I. 1959. Glaze, its meteorology and climatology, geographical distribu-


tion and economic effects. Technical Rep. EP-105. Natick, MA: US Army,
Quartermaster Research and Engineering Center, Environmental Protec-
tion Research Division.
Bernstein, B. C., and B. G. Brown. 1997. “A climatology of supercooled
large drop conditions based upon surface observations and pilot reports
of icing.” In Proc., 7th Conf. on Aviation, Range and Aerospace Meteorology,
Long Beach, California.
Birjulin, A. P., V. V. Burgsdorf, and B. J. Makhlin. 1960. Wind loads on over-
head lines. Paper 225. Paris: International Council on Large Electrical
Systems (CIGRÉ).
Bocchieri, J. R. 1980. “The objective use of upper air soundings to specify
precipitation type.” Mon. Weather Rev. 108: 596–603. https://doi.org/10
.1175/1520-0493(1980)108<0596:TOUOUA>2.0.CO;2.
Boddy, D. M., and J. Rice. 2009. “Impact of alternative galloping criteria on
transmission line design.” In Proc., ASCE Electrical Transmission and Sub-
station Structures 2009, Fort Worth, Texas, 499–510.
Brekke, G. N. 1959. Wind pressures in various areas of the United States. Cata-
log No. C13.29. Washington, DC: US Weather Bureau, National Bureau
of Standards.
Britter, R. E., J. C. R. Hunt, and K. J. Richards. 1981. “Air flow over a two-
dimensional hill: Studies of velocity speed-up, roughness effects and
turbulence.” Q. J. R. Meteorolog. Soc. 107 (451): 91–110. https://doi.
org/10.1002/qj.49710745106.
Brokenshire, R. E. 1979. Experimental study of the loads imposed on welded steel
support structures by galloping 345-kV bundled conductors. Paper A 79
551-3. New York: IEEE.
Bryant, P, 2012. Limiting the Effects of Longitudinal Loads on Small Angle
Lattice Transmission Towers, Philip Bryant, CenterPoint Energy, 2012.
BSI (British Standards Institution). 1972. Basic data for the design of build-
ings—Chapter V: Loading—Part 2: Wind Loads. BSI CP3. London: BSI.
Carper, K. L., ed. (1986). Forensic engineering: learning from failures. New
York: ASCE.
Castanheta, M. N. 1970. Dynamic behavior of overhead power lines subject to
the action of the wind. Paper No. 22-08. Paris: International Council on
Large Electrical Systems (CIGRÉ).
CEA (Canadian Electrical Association). 1998. Validation of ice accretion mod-
els for freezing precipitation using field data. 331 T 992 (A-D). Montreal:
CEA.
CENELEC (European Committee for Electrotechnical Standardization).
2001. Overhead electrical lines exceeding AC 1 kV - Part 2-20: National Nor-
mative Aspects (NNA) for ESTONIA. EN 50341. Brussels: CENELEC.
CENELEC. 2012. Overhead electrical lines exceeding AC 1 kV. General require-
ments. Common specifications. EN 50341. Brussels: CENELEC.
290 References

Chaîné, P. M., and G. Castonguay. 1974. New approach to radial ice thickness
concept applied to bundle-like conductors. Industrial Meteorology—Study
IV. Toronto: Environment Canada.
CIGRÉ (International Council on Large Electrical Systems). 2005. Overhead
conductor safe design tension with respect to Aeolian vibrations. Task Force
B2.11.04. Paris: CIGRÉ.
CIGRÉ. 2007. State of the art conductor galloping. Task Force B2.11.06. Paris:
CIGRÉ.
CIGRÉ. 2009. Overhead line design guidelines for mitigation of severe wind storm
damage. Scientific Committee B2 on Overhead Lines, B2.06.09. Paris:
CIGRÉ.
Clough, P. W., and P. Penzien. 1975. Dynamics of structures. New York:
McGraw-Hill.
Cluts, S., and A. Angelos. 1977. “Unbalanced forces on tangent transmis-
sion structures.” In Proc., IEEE Winter Power Meeting, New York, Paper
No. A77-220-7. New York: IEEE.
Colbeck, S. C., and S. F. Ackley. 1982. “Mechanisms for ice bonding in wet
snow accretions on power lines.” In Proc., 1st Int. Workshop on Atmo-
spheric Icing of Structures, Hanover, New Hampshire, 25–30.
Comellini, E., and C. Manuzio. 1968. Rational determination of design loadings
for overhead line towers. Paper 23-08. Paris: International Council on Large
Electrical Systems (CIGRÉ).
Coppin, P. A., E. F. Bradley, and J. J. Finnigan. 1994. “Measurements of flow
over an elongated ridge and its thermal stability dependence: The mean
field.” Boundary Layer Meteorol. 69 (1994): 173–199. https://doi.
org/10.1007/BF00713302.
Davenport, A. G. 1960. Wind loads on structures. Technical Paper No. 88 of
the Division of Building Research. Ottawa: National Research Council
of Canada.
Davenport, A. G. 1962. “The response of slender line-like structures to a
gusty wind.” In Proc., Inst. Civ. Eng. 23 (3): 389–408. https://doi.
org/10.1680/iicep.1962.10876.
Davenport, A. G. 1967. “Gust loading factors.” J. Struct. Div. 93 (3): 11–34.
Davenport, A. G. 1977. “The prediction of the response of structures to
gusty wind.” In Proc., Int. Research Seminar on the Safety of Structures
under Dynamic Loading, Trondheim, Norway, Vol. 1, 257–284.
Davenport, A. G. 1979. “Gust response factors for transmission line load-
ing.” In Proc., 5th Int. Conf. on Wind Engineering (IAWE), Fort Collins, CO,
899–909.
Davison et al. 1961. Alcoa overhead engineering data no. 4. Pittsburgh: Alcoa.
Den Hartog, J. P. 1932. “Transmission line vibration due to sleet.” Trans.
AIEE 51 (4): 1074–1086. https://doi.org/10.1109/T-AIEE.1932.5056223.
Durst, C. S. 1960. “Wind speeds over short periods of time.” Meteorol. Mag.
89 (1056): 181–186.
References 291

ESDU (1993). “Strong winds in the atmospheric boundary layer. Part 2:


Discrete gust speeds.” ESDU Data Item No. 83045
Elawady, A., and A. A. El Damatty. 2018. “Critical load cases for lattice
transmission line structures subjected to downbursts: Economic implica-
tions for design of transmission lines.” Eng. Struct. 159: 213–226.
El Damatty, A. A., A. Hamada, and A. Elawady. 2013. “Development of criti-
cal load cases simulating the effect of downbursts and tornadoes on trans-
mission line structures.” In Proc., 8th Asia-Pacific Conf. on Wind Engineering,
Chennai, India. https://doi.org/10.3850/978-981-07-8012-8_Key-01.
El Damatty, A. A., M. Hamada, and A. Hamada. 2015. “Simplified load
cases for lattice transmission line structures under tornadoes.” In Proc.,
14th Int. Conf. on Wind Engineering, Porto Alegre, Brazil.
Engleman, W. C., and D. Marihugh. 1970. “Forces on conductors at high
wind velocities.” Transmission and Distribution, October.
EPRI (Electric Power Research Institute). 1978. Longitudinal unbalanced loads
on transmission line structures. Rep. EL-643. Palo Alto, CA: EPRI.
EPRI. 1979. Transmission line reference book. Palo Alto, CA: EPRI.
EPRI. (1983). “Longitudinal unbalanced loads on transmission structures.”,
Computer Programs Documentation, BROD12 and BROFLX, Project
561-1 Report EL-2943-CC7.
ESDU (Engineering Science Data Unit). 1982. Strong winds in the atmospheric
boundary layer. Part 1: Mean hourly wind speed. ESDU 82026. London:
ESDU.
ESDU. 1983. Strong winds in the atmospheric boundary layer. Part 2: Discrete
gust speeds. ESDU 83045. London: ESDU.
Farr, H. H. 1980. Transmission line design manual. Denver: US Dept. of the
Interior, Water, and Power Resources Service.
Felin, B. 1988. “Freezing rain in Quebec: Field observations compared to
model estimations.” In Proc., 4th Int. Workshop on Atmospheric Icing of
Structures, Paris, 119–123.
Finnigan, J. J., M. R. Raupach, E. F. Bradley, and G. K. Aldis. 1990. “A wind
tunnel study of turbulent flow over a two-dimensional ridge.” Boundary
Layer Meteorol. 50 (1990): 277–317. https://doi.org/10.1007/BF00120527.
Finstad, K., R. Hopkinson, and K. Jones. 2009. Wet snow modeling for
Alberta—2009. Rep. for Alberta Electric System Operator (AESO).
Ottawa: VP Enterprises.
Finstad, K., and E. Lozowski. 1988. “A computational investigation of
water droplet trajectories.” J. Atmos. Oceanic Technol. 5: 160–170. https://
doi.org/10.1175/1520-0426(1988)005<0160:ACIOWD>2.0.CO;2.
Frandsen, A. G., and P. H. Juul. 1976. Cascade of tower collapses: Design crite-
ria. Paper 22-10. Paris: International Council on Large Electrical Systems
(CIGRÉ).
Fujita, T. T. 1985. The downburst: Microburst and microburst. SMRP Research
Paper 210. Chicago: Univ. of Chicago.
292 References

Fujita, T. T., and A. D. Pearson. 1973. “Results of FPP classification of 1971


and 1972 tornadoes.” In Proc., 8th Conf. on Severe Local Storms, Denver,
142–145.
Gibbon, R. R. 1984. Damage to overhead lines caused by conductor galloping.
Art. ELT_094_4. Paris: International Council on Large Electrical Systems
(CIGRÉ).
Gland, H., and P. Admirat. 1986. “Meteorological conditions for wet snow
occurrence in France—Calculated and measured results in a recent case
study on 5 March 1985.” In Proc., 3rd Int. Workshop on Atmospheric Icing
of Structures, Vancouver, 91–96.
Gong, W., and A. Ibbetson. 1989. “A wind tunnel study of turbulent flow
over model hills.” Boundary-Layer Meteorol. 49 (1989): 113–148. https://
doi.org/10.1007/BF00116408.
Goodwin, E. J., J. D. Mozer, A. M. DiGioia Jr., and B. A. Power. 1983. “Pre-
dicting ice and snow loads for transmission line design.” In Proc, 3rd Int.
Workshop on Atmospheric Icing of Structures, Vancouver, 267–275.
Gouze, S. C., and M. C. Richmond. 1982a. Meteorological evaluation of the
proposed Alaska transmission line routes. Altadena, CA: Meteorology
Research, Inc.
Gouze, S. C., and M. C. Richmond. 1982b. Meteorological evaluation of the
proposed Palmer to Glennallen transmission line route. Altadena, CA: Meteo-
rology Research, Inc.
Griffing, K. L., and Leavengood, D. C. (1973). “Transmission line failures,
part I. Meteorological phenomena—- severe winds and icing.” In Proc.,
IEEE Winter Power Mtg., Paper No. 73 CH0816-9-PWR, 5–14. New York.
Hamada, A., and A. A. El Damatty. 2011. “Behaviour of guyed transmission
line structures under tornado wind loading.” Comput. Struct. 89 (11–12):
986–1003. https://doi.org/10.1016/j.compstruc.2011.01.015.
Hamada, A., and A. A. El Damatty. 2015. “Failure analysis of guyed trans-
mission lines during F2 tornado event.” Eng. Struct. 85 (February 15):
11–25. https://doi.org/10.1016/j.engstruct.2014.11.045.
Hamada, A., A. A. El Damatty, H. Hangan, and A. Y. Shehata. 2010. “Finite
element modelling of transmission line structures under tornado wind
loading.” Wind Struct. 13 (5): 451–469.
Hangan, H., and J.-D. Kim. 2008. “Swirl ratio effects on tornado vortices in
relation to the Fujita scale.” Wind Struct. 11 (4): 291–302.
Hangan, H., D. Roberts, Z. Xu, and J. Kim. 2003. “Downburst simulation.
Experimental and numerical challenges.” In Proc., 11th Int. Conf. on Wind
Engineering, Lubbock, Texas, June 1–5.
Havard, D. G., and J. C. Pohlman. 1980. Control of galloping conductors by
detuning. Paper 22-05. Paris: International Council on Large Electrical
Systems (CIGRÉ).
References 293

Havard, D. G., A. S. Paulson, and J. C. Pohlman. 1982. The economic benefits


of controls for conductor galloping. Paper 22-02. Paris: International Coun-
cil on Large Electrical Systems (CIGRÉ).
Hjelmfelt, M. R. 1988. “Structure and life cycle of microburst outflows
observed in Colorado.” J. App. Meteorol. 27 (8): 900–927. https://doi.org
/10.1175/1520-0450(1988)027<0900:SALCOM>2.0.CO;2.
Hoerner, S. F. 1958. Fluid-dynamic drag. Vancouver, WA: Hoerner Fluid
Dynamics.
Holmes, J. D. 2001. Wind loading of structures. New York: Spon.
Holmes, J. D., H. M. Hangan, J. L. Schroeder, C. W. Letchford, and K. D.
Orwig. 2008. “A forensic study of the Lubbock-Reese downdraft of 2002.”
Wind Struct. 11 (2): 137–152. https://doi.org/10.12989/was.2008.11.2.137.
Hoskings, J. R. M., and J. R. Wallis. 1987. “Parameter and quantile estima-
tion for the generalized Pareto distribution.” Technometrics 29 (3): 339–
349. https://doi.org/10.2307/1269343.
Houghton, E. L., and N. B. Carruthers. 1976. Wind forces on building and
structures. New York: Wiley.
IEC (International Electrotechnical Commission). 2003a. Design criteria of
overhead transmission lines. IEC 60826, Ed. 3.0. Geneva: IEC.
IEC. 2003b. Loading and strength of overhead transmission lines. IEC 60826, 3rd
ed. Geneva: IEC.
IEC. 2017. Design criteria of overhead transmission lines. IEC 60826, Ed. 4.0.
Geneva: IEC.
IEEE. 1999. “Limitations of the ruling span method for overhead line con-
ductors at high operating temperatures.” IEEE Trans. Power Deliv. 14 (2):
549–560. https://doi.org/10.1109/61.754102.
IEEE. 2003. IEEE guide to the installation of overhead transmission line conduc-
tors. IEEE 524. New York: IEEE.
IEEE. 2004. IEEE standard for fall protection for utility work. IEEE 1307. New
York: IEEE.
IEEE. 2012. National electrical safety code. ANSI C2. New York: IEEE.
IEEE. 2017. National electrical safety code. ANSI C2-2017. New York: IEEE.
ISO (International Standards Organization). 1999. Atmospheric icing of struc-
tures. ISO 12494. Geneva: ISO.
James, W. D. 1976. “Effects of Reynolds number and corner radius on
two-dimensional flows around octagonal, dodecagonal, and hexdecago-
nal cylinders.” Ph.D. diss., College of Engineering, Iowa State Univ.
Janney, J. R. (1979). Guide to investigation of structural failures., New York:
ASCE.
Jones, K. 1996. “A simple model for freezing rain loads.” In Proc., 7th Int.
Workshop on Atmospheric Icing of Structures, Chicoutimi, Quebec,
412–416.
Jones, K. F. 1998. “A simple model for freezing rain ice loads.” Atmos. Res.
46 (1–2): 87–97. https://doi.org/10.1016/S0169-8095(97)00053-7.
294 References

Jung, S., and F. J. Masters. 2013. “Characterization of open and suburban


boundary layer wind turbulence in 2008 Hurricane Ike.” Wind Struct. 17
(2): 135–162. https://doi.org/10.12989/was.2013.17.2.135.
Kempner, L., Jr. 1997. “Longitudinal impact loading on electrical transmis-
sion line towers—A scale model study.” Ph.D. diss., Dept. of Civil and
Environmental Engineering, Portland State Univ.
Keselman, L. M., and Y. Motlis. 1996. “Enhanced analytical design method
for overhead line conductors in non-level spans.” In Proc., IEEE/PES
Transmission and Distribution Conf., Los Angeles, 359–366.
Keselman, L. M., and Y. Motlis. 1998. “Application of the ruling span con-
cept for overhead lines in mountainous terrain.” IEEE Trans. Power Deliv.
13 (4): 1385–1390. https://doi.org/10.1109/61.714512.
Kuroiwa, D. 1962. A study of ice sintering. Research Rep. 86. Hanover, NH:
US Army Cold Regions Research and Engineering Laboratory.
Kuroiwa, D. 1965. Icing and snow accretion on electric wires. Research Report
123. Hanover, NH: US Army Cold Regions Research and Engineering
Laboratory.
Langmuir, I., and K. Blodgett. 1946. “Mathematical investigation of water
droplet trajectories.” US Army Air Forces Tech. Rep. 5418. Washington,
DC: Office of the Publication Board, Dept. of Commerce.
MacDonald, A. 1975. Wind loading on buildings. New York: Wiley.
Makkonen, L. 1996. “Modeling power line icing in freezing precipitation.”
In Proc., 7th Int. Workshop on Atmospheric Icing of Structures, Chicoutimi,
Quebec, 195–200.
Mara, T. G. 2014. “Influence of solid area distribution on the drag of a
two-dimensional lattice frame.” J. Eng. Mech. 140 (3): 644–649. https://
doi.org/10.1061/(ASCE)EM.1943-7889.0000681.
Mara, T. G. 2015. “Updated gust response factors for transmission line
loading.” In Proc., Electrical Transmission and Substation Structures Conf.
2015, Branson, Missouri. https://doi.org/10.1061/9780784479414.037.
Mara, T. G., J. K. Galsworthy, and E. Savory. 2010. “Assessment of vertical
wind loads on lattice framework with application to thunderstorm winds.”
Wind Struct. 13 (5): 413–431. https://doi.org/10.12989/was.2010.13.5.413.
Mathur, R. K., A. H. Shah, P. G. S. Trainor, and N. Popplewell. 1986. Dynam-
ics of guyed transmission tower system. Paper 86 SM 414-7. New York:
IEEE.
McComber, P., R. Martin, G. Morin, and L. V. Van. 1982. “Estimation of
combined ice and wind load on overhead transmission lines.” In Proc.,
1st Int. Workshop on Atmospheric Icing of Structures, Hanover, New Hamp-
shire, 143–153.
McCormick, T., and J. C. Pohlman. 1993. “Study of compact 220 kV line
system indicates need for micro-scale meteorological information.” In
Proc., 6th Int. Workshop on Atmospheric Icing of Structures, Budapest.
References 295

McDonald, J. R. 1983. A methodology for tornado hazard probability assessment.


NUREG/CR-3058. Washington, DC: Nuclear Regulatory Commission.
Mehta, K. C., and T. Lou. 1983. Force coefficients for transmission towers. Elec-
tric Power Research Institute (EPRI) Rep., Technical Agreement TPS
82–623. Washington, DC: EPRI.
Mehta, K. C., J. E. Minor, and J. R. McDonald. 1976. “Windspeeds analyses
of April 3–4, 1974 tornadoes.” J. Struct. Div. 102 (9): 1709–1724.
Meteorological Research Institute. 1977. Ontario Hydro wind and ice loading
model. MRI 77 FR-1496. Toronto: Ontario Hydro.
Miller, C. A. 2011. “Revisiting the Durst gust factor curve.” Can. J. Civ. Eng.
38 (9): 998–1001.
Miller, C. A., and A. G. Davenport. 1998. “Guidelines for the calculation of
wind speed-ups in complex terrain.” J. Wind Eng. Ind. Aerodyn. 74–76
(April 1): 189–197. https://doi.org/10.1016/S0167-6105(98)00016-6.
Minor, J. E., J. R. McDonald, and K. C. Mehta. 1977. The tornado: An
engineering-oriented perspective. NOAA Technical Memorandum ERL
NSSL-82. Norman, OK: National Severe Storms Laboratory.
Modi, V. J., and J. E. Slater. 1983. “Unsteady aerodynamics and vortex-
induced aeroelastic instability of a structural angle section.” J. Wind Eng.
Ind. Aerodyn. 11 (1–3): 321–334.
Mozer, J. D., J. C. Pohlman, and J. F. Fleming. 1977. “Longitudinal load
analysis of transmission line system.” IEEE Trans. Power Apparatus Syst.
96 (5): 1657–1665. https://doi.org/10.1109/T-PAS.1977.32495.
Mozer, J. D., and R. J. West. 1983. “Analysis of 500 kV tower failures.” In
Proc., Pennsylvania Electric Association, Norristown, Pennsylvania.
National Bureau of Standards. 1920. National electrical safety code. Handbook
No. 3. Washington, DC: US Dept. of Commerce.
Nebraska Public Power District. 1976. “The storm of March 29, 1976.”
Columbus, NE: Nebraska Public Power District.
Nigol, O., and Havard, D. G. 1978. Control of torsionally induced conductorsby
detuning. Paper A 78 125-7. Piscataway, NJ: IEEE.
NOAA (National Oceanic and Atmospheric Administration). (1959–Pres-
ent). Storm data. Accessed June 10, 2020. https://www.ncdc.noaa.gov/
stormevents/.
NRCC (National Research Council of Canada). 2010. National building code
of Canada (NBCC). Ottawa: NRCC.
Orwig, K. D. and J. L. Schroeder. 2007. “Near-surface wind characteristics
of severe thunderstorm outflows.” J. Wind Eng. Ind. Aerodyn. 95 (7):
565–584. https://doi.org/10.1016/j.jweia.2006.12.002.
Ostendorp, M. 1997. Longitudinal load and cascading failure risk assessment
(CASE). Vol. I–IV of EPRI Rep. TR-107087. Palo Alto, CA: Electric Power
Research Institute (EPRI).
Paz, M. 1980. Structure dynamics—Theory and computations. New York: Van
Nostrand Reinhold.
296 References

Peabody, A. B. 1993. “Snow loads on transmission and distribution lines in


Alaska.” In Proc., Sixth Int. Workshop on Atmospheric Icing of Structures,
Budapest.
Peterka, J. A. 1992. “Improved extreme wind prediction for the United
States.” J. Wind Eng. Ind. Aerodyn. 41 (1–3): 533–541. https://doi.org/10
.1016/0167-6105(92)90459-N.
Peterka, J. A., K. Finstad, and A. K. Pandy. 1996. Snow and wind loads for Tyee
transmission line. Fort Collins, CO: Cermak Peterka Petersen.
Peyrot, A. H. 1985. “Microcomputer-based nonlinear structural analysis of
transmission line systems.” IEEE Trans. Power Apparatus Syst. PAS-104
(11): 3236–3244. https://doi.org/10.1109/TPAS.1985.318837.
Peyrot, A. H. and H. J. Dagher. 1984. “Reliability-based design of transmis-
sion lines.” J. Struct. Eng. 110 (11): 2758–2777. https://doi.org/10.1061/
(ASCE)0733-9445(1984)110:11(2758).
Pintar, A. L., E. Simiu, F. T. Lombardo, and M. Levitan. 2015. Maps of
non-hurricane non-tornadic wind speeds with specified mean recurrence inter-
vals for the contiguous United States using a two-dimensional Poisson process
extreme value model and local regression. NIST Special Publication 500-301.
Gaithersburg, MD: NIST.
Pohlman, J. C., and D. Havard. 1979. Field research on the galloping of iced
conductors—A status report. Paper A 79 551-3. New York: IEEE.
Pohlman, J. C., and C. B. Rawlins. 1979. “Field testing an anti-galloping
concept on overhead lines.” In Proc., 7th IEEE/PES Transmission and Dis-
tribution Conference and Exposition, Atlanta. https://doi.org/10.1109/
TDC.1979.712640.
Rawlins, C. B. 1981. “Analysis of conductor galloping field observations
single conductors.” IEEE Trans. Power Apparatus Syst. PAS-100 (8): 3744–
3751. https://doi.org/10.1109/TPAS.1981.317017.
REA (Rural Electrification Association). 1980. “Design manual for high
voltage transmission lines,” Bulletin 62-I. Washington, DC: US Dept. of
Agriculture, Rural Electrification Administration.
Richards, D. J. W. 1965. “Aerodynamic properties of the Severn crossing
conductor.” In Proc., Symp. 16, Middlesex, United Kingdom, Paper 8.
Richardson, A. S. 1983. The wind damper method of galloping control. DOE/
CE/15102T1. Springfield, VA: National Technical Service, US Dept. of
Commerce.
Richardson, A. S. 1986. Longitudinal dynamic loading of steel pole transmission
lines. Paper WM 189-5. New York: IEEE.
Richmond, M. C. 1985. Meteorological evaluation of Bradley Lake hydroelectric
project 115 kV transmission line route. Torrance, CA: M. C. Richmond
Meteorological Consulting.
Richmond, M. C. 1991. Meteorological evaluation of Tyee Lake hydroelectric
project transmission line route, Wrangell to Petersburg. Torrance, CA: M.C.
Richmond Meteorological Consulting.
References 297

Richmond, M. C. 1992. Meteorological evaluation of Tyee Lake hydroelectric


project transmission line route, Tyee power plant to Wrangell. Torrance, CA:
M.C. Richmond Meteorological Consulting.
Riley, M. J., L. Kempner Jr., W. H. Mueller III, and G. T. Gobo. 2002. “A
comparison of seismic (dynamic) and static load cases for lattice electric
transmission towers.” In Proc., ASCE Electrical Transmission in a New Age,
Omaha, Nebraska, 420–426. https://doi.org/10.1061/40642(253)39.
Robbins, C. C., and J. V. Cortinas Jr. 1996. “A climatology of freezing rain
in the contiguous United States: Preliminary results.” In Preprints, 15th
AMS Conf. on Weather Analysis and Forecasting, American Meteorological
Society, Norfolk, Virginia.
Sachs, P. 1972. Wind forces in engineering. New York: Pergamon.
Sachs, P. 1978. Wind forces in engineering, 2nd ed. New York: Pergamon.
Sakamoto, Y., K. Mizushima, and S. Kawanishi. 1990. “Dry snow type
accretion on overhead wires: growing mechanism, meteorological con-
ditions under which it occurs and effect on power lines.” In Proc., 5th
Int. Workshop on Atmospheric Icing of Structures, Tokyo, Paper 5-9.
Scruton, C. and C. W. Newberry. 1963. “On the estimation of wind loads
on buildings and structural design.” In Proc., Inst. Civil Eng. 25 (6):
97–126. https://doi.org/10.1680/iicep.1963.10660.
Shan, L., and L. Marr. 1996. Ice storm data base and ice severity maps. Palo
Alto, CA: Electric Power Research Institute (EPRI).
Shehata, A. Y., and A. A. El Damatty. 2007. “Behaviour of guyed transmis-
sion line structures under downburst wind loading.” Wind Struct. 10 (3):
249–268. https://doi.org/10.12989/was.2007.10.3.249.
SIA (Swiss Association of Engineers and Architects). 1956. Standards for load
assumptions, acceptance and inspection of structures. Tech. Paper No. 160.
Zurich: SIA.
Simiu, E., and R. H. Scanlan. 1996. Wind effects on structures: Fundamentals
and applications to design, 3rd ed. New York: Wiley.
Slater, J. E., and V. J. Modi. 1971. “On the wind-induced vibration of struc-
tural angle sections.” In Proc., IAWE, Third Int. Conf. on Wind Effects on
Building and Structures, Tokyo.
Snyder, W. H., and R. E. Britter. 1987. “A wind tunnel study of the flow
structure and dispersion from sources upwind of three-dimensional
hills, atmospheric environment.” Atmos. Environ. 21 (4): 735–751.
https://doi.org/10.1016/0004-6981(87)90070-9.
Stallabrass, J. R. and P. F. Hearty. 1967. The icing of cylinders in conditions of
simulated freezing sea spray. Mechanical Engineering Rep. MD-50, NRC
No. 9782. Ottawa: National Research Council.
Tattelman, P., and I. Gringorten. 1973. Estimated glaze ice and wind loads at
the earth’s surface for the contiguous United States. Rep. AFCRL-TR-73-0646.
Bedford, MA: US Air Force Cambridge Research Laboratories.
298 References

Taylor, P. A., P. J. Mason, and E. F. Bradley. 1987. “Boundary layer flow over
low hills, a review.” Boundary Layer Meteorol. 39 (1–2): 107–132. https://
doi.org/10.1007/BF00121870.
Tecson, J. J., T. T. Fujita, and R. F. Abbey Jr. 1979. “Statistics of U. S. torna-
does based on the DAPPLE tornado tape.” In Proc., 11th Conf. on Severe
Local Storms, Kansas City, Missouri, 227–234.
Thayer, E. S. 1924. “Computing tensions in transmission lines.” Electr.
World 84 (2): 72–73.
Thomas, M. B. 1981. “Broken conductor loads on transmission line struc-
tures,” Ph.D. diss., College of Civil and Environmental Engineering,
Univ. of Wisconsin–Madison.
Thrasher, W. J. 1984. “Halt redundant member failure of 765-KV towers.”
Trans. Distrib., May.
Trainor, P. G. S., N. Popplewell, A. H. Shah, and C. K. Wong. 1984. Estima-
tion of fundamental dynamic characteristics of transmission towers. Paper 84
SM 525-2. New York: IEEE.
Twisdale, L. A. 1982. “Wind-loading underestimate in transmission line
design.” Trans. Distrib., December, 40–46.
USFS (United States Forest Service). 1994. Telecommunications handbook, R3
supplement 6609.14-94-2. Forest Service Handbook, FSH 6609, 14, Wash-
ington, DC: USFS.
Vellozzi, J. W., and E. Cohen. 1968. “Gust response factors.” J. Struct. Div.
94 (6): 1295–1314.
Vickery, P. J., and P. F. Skerlj. 2005. “Hurricane gust factors revisited.” J.
Struct. Eng. 131 (5): 825–832. https://doi.org/10.1061/(ASCE)0733-944
5(2005)131:5(825).
Vickery, P. J., D. Wadhera, J. Galsworthy, J. A. Peterka, P. A. Irwin, and L.
A. Griffis. 2010. “Ultimate wind load design gust wind speeds in the
United States for use in ASCE-7.” J. Struct. Eng. 136 (5): 613–625. https://
doi.org/10.1061/(ASCE)ST.1943-541X.0000145.
Walmsley, J. L., P. A. Taylor, and T. Keith. 1986. “A simple model of neu-
trally stratified boundary-layer flow over complex terrain with surface
roughness modulations (MS3DJH/3R).” Boundary-Layer Meteorol. 36
(1986): 157–186. https://doi.org/10.1007/BF00117466.
Wang, Q. J. 1991. “The POT model described by the generalized Pareto
distribution with Poisson arrival rate.” J. Hydrol. 129 (1–4): 263–280.
https://doi.org/10.1016/0022-1694(91)90054-L.
Wardlaw, R. L. 1967. “Wind-excited vibrations of slender beams with angle
cross-sections.” In Proc., Int. Research Seminar—Wind Effects on Buildings
and Structures, Ottawa.
Watson, L. T. 1955. Drag of bare stranded cables. Tech. Memorandum 114.
Melbourne, AU: Aeronautical Research Laboratories.
Wen, Y. K., and S.-L. Chu. 1973. “Tornado risks and design wind speed.” J.
Struct. Div. 99 (12): 2409–2421.
References 299

Whapham, R. 1982. Field research for control of galloping with air flow spoilers.
Cleveland: Preformed Line Products.
Whitbread, R. S. 1979. “The influence of shielding on the wind forces expe-
rienced by arrays of lattice frames.” In Proc., 5th Int. Conf. on Wind Engi-
neering (IAWE), Fort Collins, Colorado, 405–420.
White, H. B. 1979. Some destructive mechanisms activated by galloping conduc-
tors. Report No. A79 106-6. New York: IEEE.
White, H. B. 1999. “Galloping of ice covered wires of a transmission line.”
In Proc., ASCE Cold Regions Conf., Lincoln, New Hampshire, 790–798.
Winkelman, P. F. 1959. Sag-tension computations and field measurements of
Bonneville Power Administration. AIEE Paper 59-900. New York: Ameri-
can Institute of Electrical Engineers.
Yip, T.-C. 1995. “Estimating icing amounts caused by freezing precipitation
in Canada.” Atmos. Res. 36 (3–4): 221–232. https://doi.org/10.101
6/0169-8095(94)00037-E.
Yip, T. C., and P. Mitten. 1991. Comparisons between different ice accretion
models. Ottawa: Canadian Electrical Association.
Young, W. R. 1978. “Freezing precipitation in the southeastern United
States.” M.S. diss., Dept. of Atmospheric Sciences, Texas A&M Univ.

ADDITIONAL RESOURCES

Bryant, P. J. 2012. “Limiting the effects of longitudinal loads on small angle


lattice transmission towers.” In Proc., ASCE Electrical Transmission and
Substation Structures 2012, Columbus, Ohio, 205–216. https://doi.
org/10.1061/9780784412657.018.
Elawady, A., and A. A. El Damatty. 2016. “Longitudinal force on transmission
structures due to non-symmetric downburst conductor loads.” Eng. Struct.
127 (Nov): 206–226. https://doi.org/10.1016/j.engstruct.2016.08.030.
Freimark, B., et al. 2012. “The effect of broken wire loads on EHV transmis-
sion structure design.” In Proc., ASCE Electrical Transmission and Substa-
tion Structures Conf. 2012, Columbus, OH, 166–182. https://doi.
org/10.1061/9780784412657.015.
Mara, T. G., and R. H. Behncke. 2015. “Examination of yawed wind loading on
transmission towers.” In Proc., Electrical Transmission and Substation Struc-
tures Conf. 2015, Branson, MO. https://doi.org/10.1061/9780784479414.042.
Peabody, A. B. 2004. “Applying shock damping to the problem of transmis-
sion line cascades.” Ph.D. thesis, Dept. of Civil Engineering and Applied
Mechanics, McGill Univ.
INDEX

acceleration and flow  43 ANSI C29  9


accretion of ice  56–57 ANSI C119  9
on aerial marker balls  66 ANSI Standard O5
maps 58–66 5.1 9
models 57 5.2 9
wind load relation  66 5.3 9
ACI 318, Building Code ASCE 7, Minimum Design Loads and
Requirements for Structural Associated Criteria for
Concrete  9 Buildings and Other
aerodynamic damping  33 Structures  22, 31
air density coefficient. 7-05 13
See coefficient of air density
7-10 13
AISC 360, Specification for Structural
7-16 10, 13, 19, 21–22, 24–26, 38,
Steel Buildings  9
44–45
alignment convergence  98
ASCE 10, Design of Latticed Steel
analysis
Transmission Structures 9
computational 90–91
consecutive tension section  90 ASCE 48, Design of Steel
extreme value  65 Transmission Pole Structures  9
single tension section  90 ASCE 104, Recommended Practice for
angle Fiber-Reinforced Polymer
centerline 98 Products for Overhead Utility
line 90, 94 Line Structures  9
vertical 86 ASCE Manual of Practice No. 111,
yaw 36–38, 42 Reliability-based Design of
ANSI/ASSE Z359.1 Fall Protection Utility Pole Structures  7, 8
Code  77 ASCE Manual of Practice No. 123,
ANSI C2 National Electric Safety Prestressed Concrete
Code  77 Transmission Pole Structures  9

301
302 INDEX

ASCE Manual of Practice No. 141, component, resonant  31, 33


Wood Pole Structures for condition
Electrical Transmission Lines  9 microclimatic 6
ASCE Technical Council on site-specific 69
Lifeline Earthquake condition of wire  86
Engineering 80 “final after creep”  87–89
atmospheric boundary layer  25, “final after load”  87–89
28, 31 “initial” 86
flow over hills  45 conductor displacement,
B horizontal. See blow-out
Behncke, R.H.  31 construction loads  15, 71
blow-out 96–97 containment, failure  51. See
deviations from the also failure containment
orthogonal 97 methods
example calculation  110–112 containment philosophies,
shifted weight span  96 failure 71
boundary layer. See atmospheric corner radius ratio  44
boundary layer corrosion 88
broken wire loads  70
critical facilities  4
C
D
catenary 93
damping
constant 88, 94
aerodynamic 33
curve 95
structural 77
hyperbolic equations  93
data
traditional constant  113–114
historical 18, 56
channeling, wind  45
CIGRÉ Task Force B2.11.04  88 ice 57
clashing 78–79 regional wind  24
climate change  18 Davenport, A.G.  30, 35
climbing devices  77 dead-end points  79, 86
coefficient of deflection 9–10, 76
variation of the strength  8 design loads  45, 97
wind pressure exposure  20, 25, devices for climbing  77
28–29, 30 downbursts 49–51
coefficient of air density  20–21 velocity. See velocity
nominal value  21 associated wind field  50–51
coefficient of force  34–40 E
aspect ratio  36 Earthquake Engineering Research
for conductors and ground Institute 80
wires 37 earthquake loads  16, 80–81
for ice  64 electrical flashovers  88
for latticed truss structures  38 elongation, wire  79, 86–87
for pole structures  43–44 equations, gust response. See gust
solidity ratio  36 response factor
INDEX 303

equilibrium configuration of inertia 35


wire 93 vertical 96–97
events viscous 35
convective 10 wind 20–21
cyclical loading  17 force coefficient. See coefficient of
examples, wind calculation. force
See wind calculation foundation, differential settlement
examples of 82
Exposure Categories  25–28 F-scale 52
Exposure B  25–26 Fujita, T.T.  49, 52
Exposure C  22, 26, 31, 58 G
Exposure D  26 galloping 78–80
F flashovers 78
facilities, critical  4 load increase  79
factor, gust response. See gust mitigation 80
response factor sag 79
parameter removals  31 geotechnical hazard assessment  82
for structure  33–34 The Great Lakes  26
failure gust factors  30–31
acceptable levels of  4–5 gust response factor  20, 30–32
cascading 15, 69, 70–71 of Davenport  30
component 70 equations 31
containment 51. See also failure for wires  33
containment methods H
containment philosophies  71 hardware fatigue  15, 17, 88
designing for  5 harmonic resonance  77
mitigation 5 hazards
predicted probabilities  14 hazard assessment,
statistical probability of  7. See geotechnical 82
also load and resistance factor seismic. See seismic hazards
design (LRFD) height
failure containment methods effective 28–30
Bonnevile Power gradient 28
Administration (BPA)  71 Ho, E.  31
Broken Wire Load (BWL)  71 Hoerner, S.F.  35
Electric Power Research hurricane. See also winds,
Institute (EPRI)  71 high-intensity
Residual Static Load (RSL)  71, 86 coastline 26
fault zone  81 prone regions  13
flashovers 78 I
electrical 88 ice 56
flexibility, longitudinal  86, 89–90 data 57
flow acceleration  43 force coefficient  64
force glaze 57
304 INDEX

ice (continued) exclusion 8


hoarfrost 57 failure 9
in-cloud icing  57 lifting 71
precipitation 56 tension 88–89
thickness 58 linear exposure  6
topographic factor of line failure  49
thickness 60 line security loads  70
unbalanced case  70 liquefaction 82
ice accretion. See accretion of ice load and resistance factor design
IEC 60826  88 (LRFD) 6–8
IEEE Standard 524, “IEEE Guide to load factors  7–8
the Installation of Overhead nominal strength  7
Transmission Line strength reduction factor  7–8
Conductors” 73, 75 load factors  7–8
IEEE Standard 1307 “Fall loading of nomenclature  99
Protection for Utility loads
Work” 77 broken wire  70
impact damage,   88 construction 15, 71
inertia force  35 dead-end connection  76
insulators, suspension  70, 86, 90 design 45, 97
intensity, turbulence  31 earthquake 16, 80–81
issues, loading  85 effects 19, 76, 77
J legislated 16
James, W.D.  44 line security  70
K longitudinal 3, 15, 69–71, 86, 88
kinetic energy  21 maintenance 15, 71, 76–77
L path 71, 76
landslides 82 Residual Static Load  71, 86
latticed truss structures  38 seismic inertial  81–82
recommended force stringing block  75
coefficient 38 transverse 3, 37, 73, 86
related calculation unbalanced 70
examples 99–110 vertical 3, 73, 76, 86, 96
square-section 38 loads, weather-related  8, 10, 19–67
triangular-section 38 extreme ice  11
wind load calculation  40, 48–51 high-intensity wind  10–11,
“Wind on Face” calculation 49–55
method 40–41 unequal ice  66
“Wind on Member” calculation wind. See wind loads
method 42 loads, wire unit  91–95
lightning strikes  14, 88 Horizontal Unit Load  91
limits summed vector load  91
damage 9–10 Vertical Unit Load  91–92
deflection/serviceability 10 longitudinal flexibility  86, 89–90
INDEX 305

longitudinal loads. See loads, Pearson, A.  52


longitudinal period, return  11
Lou, T.  35 permanent ground
LRFD (load and resistance displacement 82
factor design). See load and phase transpositions  98
resistance factor design portal, guyed  49
M probability distributions  7
maintenance
transmission line  76–77 R
maps radial uniform thickness  57
contour 21 rain, freezing  57, 61, 64–65, 66
ice accretion  58–66 relation to wind load  66
linear interpolation  22 reliability 6–7
wind speed  21–24 MRIs and  13–15
Mara, T.G.  31, 38 reliability-based design  6–7
mean recurrence interval service 14
(MRI) 11–14 structural 14
and design load  14 Residual Static Load  71, 86
reliability and  13 restricted suspension points  90
Mehta, K.C.  35 Reynolds number  35, 37, 43
meteorologist, consultant  22, 45,
57, 61 S
method, ruling span  89–90 Sachs, P.  35
mitigation 80 sagging
failure 5 low point formula  93
mountains 45 projected low point  95–96
Mulhall, Oklahoma  54 wire 79, 88–89, 89
N seismic hazards  16, 81–82
National Electrical Safety Code fault zone  81
(NESC) 16, 88 landslides 82
Rule 010  16 liquefaction 81
Rule 251  37 permanent ground
National Oceanic and Atmospheric displacement 81
Administration 52 strike/slip fault  81
National Weather Service  52 separation distance  36
NESC (National Electrical Safety service redundancy  4
Code). See National Electrical shapes, member  35
Safety Code bluff 35
rounded 35, 38–39
O sharp-edged (flat-sided)  35
Occupational Safety and Health streamlined 35
Administration 77 shielding 36, 38, 43
OSHA 1910.269  77 factors 36
outages 78 solidity ratio  36
306 INDEX

slack 16, 81 support points  80


transference of  70 surface roughness  44
slope suspension 3
3H:1V 73 restricted points  90
pulling line  73 supports 90
snow, sticky  57–58 systems. See also wire systems
spans. See also weight spans fall arrest  77
adjacent 90, 92 structural support  3–4
ahead 15, 102 T
back 73, 102 tangent locations  92
characteristics 85 tension
disturbances 90 design 4
inclination 97 differential 4
length 89, 94 equalization of  89
low point calculation  92–95 imbalances 70
reduction factor  51 limits 88–89
reduction factors  51 unbalanced 4
span-specific disturbances  90 wires and  70, 73–75, 85–87
Special Wind Regions  22 termination points  4
Hawaii 22–23 time signature loads  16–17
Spencer, South Dakota  54 topographic factors  48–49
states, limit  8–10 tornadoes 51–55
Storm Prediction Center  52 F-Scale 52
strength wind field  54–55
coordinating 5 transmission pole structures  30,
longitudinal 70–71 33, 37, 46, 48, 78
nominal 7, 9 turbulence intensity  31
structure 97 V
strength reduction factor  7–8. See velocity
also strength, nominal jet velocity  51
strike/slip fault  81 velocity’s radial component  50
stringing block loads  75 velocity’s vertical
structural damping  77 component 50
Structure Coordinate System  99 Venturi effect  45
structures. See also latticed truss vertical force  96–97
structures, transmission pole vertical loads. See loads, vertical
structures vertical to delta configuration  98
angle 3 vibration
dead-end/strain 4, 70, 89 aeolian 88
H-frame 36, 43 ground-induced 80
pole 43–44 mitigation 88
tangent 3, 66 viscous force  35
uphill 76 vortex shedding  77–78
INDEX 307

weather-related loads. See loads, yawed 37–38, 43–44


weather-related wind speed
weight spans  92, 96–98 basic 21
example calculation with through canyons  48
blow-out 110–112 critical 35
temperature variation  97 over hills and
wind calculation examples  102 escarpments 45–46
with extreme radial glaze increase of  45–46
ice 106 over mountains  45
with yaw angles  104–106 power law increase with
wind channeling  45 height 58
wind field  50–51 topographic effects  44
wind force  20–21 wire. See also tension, wires and,
expression 20 wire systems
wind loads  10, 20–49 angle 98
air density coefficient, Q  21 breakage 15. See also condition
basic wind speed  21 of wire
gust response factor. See factor,
conductor 86
gust response
electrical properties  91
relation to ice  58, 64–66, 66
elongation 79, 86–87
latticed truss structure
equilibrium configuration  93
calculation 40–43
guy 76
prediction 19
motion fatigue  88
topographic effects  22, 44
physical properties  91
wind force  20–21
wind force expression  20 stringing 76
wind pressure coefficient  25, 30 unit loads. See loads, wire unit
wind pressure exposure wire attachment elevation
coefficient 20, 28–29 points 95
winds wire systems  2–3, 85–86
Chinook 45 consecutive tension section
downslope 45 analysis 90
force expression  20 longitudinal loads  3
high-intensity 49–55 single tension section
local canyon  45 analysis 90
oscillation 78 tension sections  86–87, 89
Santa Ana  45 transverse loads  3
spans 91 vertical loads  3
standing wave  45 worker weight loads  77
synoptic 10, 49 Y
variation 70 yaw angle  36–38, 42
vibration from  77–78 Z
wind tunnel tests  37, 42, 45, 48 zones, loading  6

You might also like