Download as pdf or txt
Download as pdf or txt
You are on page 1of 310

I

I Technical Report No. 2

I
CONNECTIONS FOR
I PRECAST PRESTRESSED
I CONCRETE BUILDINGS
including earthquake resistance
I
I L. D. Martin and W. J. Korkosz

I
I
I
I
I
I
I
I
I
I
Copyright 0 1982
I
I
By Prestressed Concrete Institute

All rights reserved. This book or any part thereof may not be
reproduced in any form without the’written permission of the

I
Prestressed Concrete Institute.

I
Library of Congress Catalog Card Number 82-81939
ISBN O-937040-20.7

I
I
Printed in U.S.A

I
I
I
I
I
I
I.
I
I
CONNECTIONS FOR
I PRECAST PRESTRESSED
I CONCRETE BUILDINGS
including earthquake resistance
I
I by
L. D. Martin and W. J. Korkosz

I
I
I A Research Investigation

I by
The Consulting Engineers Group, Inc.

I 1701 East Lake Avenue


Glenview, Illinois 60025

I supported by

National Science Foundation

I
Washington, D.C.

and

I Prestressed Concrete Institute


Chicago, Illinois

I
I
I
March 1982

I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
TABLE OF CONTENTS
I
1. INTRODUCTION AND OVERVIEW. . . . . . . . . . . . . . . . . . . . . . . . . . 1.1
I 1.1 Scope of This Study . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.2
1.2 Development of the Precast, Prestressed Concrete

I Industry in North America . . . . . . . . . . . . . . . . . . . . . . . . .


1.2.1
1.2.2
History . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Development of Standard Products . . . . . . . . . . . .
1.3
1.3
1.4

I
1.2.3 Inhibitors to the Growth of the Industry . . . . . . . . 1.6

References-Part1 ...................................... 1.9

I 2. RARTHQUARE AND OTHER EXTREME LOADING . . . . . . . . . . . 2.1

I
2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.1
2.2 Nature of Earthquake Forces on Buildings . . . . . . . . . . . . . 2.2
2.3 Predicting Structural Behavior . . . . . . . . . . . . . . . . . . . . . . . 2.4
2.3.1 Dynamic Analysis Procedures . . . . . . . . . . . . . . . . . 2.5

I
2.3.2 Seismic Building Design Codes . . . . . . . . . . . . . . . . 2.12
2.4 Design Philosophy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.19
2.4.1 Energy Principles . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.20

I
2.4.2 Displacement Ductility and Load Reduction
Factors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.20
2.4.3 Curvature Ductility . . . . . . . . . . . . . . . . . . . . . . . . . 2.24
2.5 Lateral Load Resisting Systems . . . . . . . . . . . . . . . . . . . . . . 2.27

I 2.5.1
2.5.2
2.5.3
Shear Wall Systems . . . . . . . . . . . . . . . . . . . . . . . . . . 2.28
Precast Frame Systems . . . . . . . . . . . . . . . . . . . . . . 2.37
Braced Frames . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.37

I
2.5.4 Floor and Roof Diaphragms . . . . . . . . . . . . . . . . . . . 2.39
2.6 Dynamic Characteristics of Precast Connections . . . . . . . 2.42
2.6.1 Shear Connections . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.42

I
2.6.2 Beam-Column Connections . . . . . . . . . . . . . . . . . . . 2.48
2.7 FutureTesting . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.54
2.8 Performance of Precast Systems in Previous
Earthquakes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.57

I
2.9 Structural Integrity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.60
2.9.1 Definitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.60
2.9.2 Philosophy of Design for Structural

I
Integrity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.61
2.9.3 Application to Severe Earthquake
Resistance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.63
2.9.4

I
Relationship to Service Load Design
Procedures.. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.63
2.9.5 Current Design Provisions for Structural
Integrity.. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.64

I
2.10 Hurricanes, Tornadoes and Other Wind Loadings . . . . . . . . 2.69
2.10.1 Comparison with Earthquake Loading . . . . . . . . . . 2.69
2.10.2 Building Code Requirements . . . . . . . . . . . . . . . . . . 2.69

I
2.10.3 Common Practice . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.70
2.10.4 Recommendations.. . . . . . . . . . . . . . . . . . . . . . . . . . 2.71

References- Part2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.74

I i

I
I
3. SELECTION AND DESIGN OF CONNECTIONS.. , . . . . . . . . . . . . 3.1
I
3.1 Criteria for Connections of Precast Concrete
Members . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.1
I
3.1.1 Strength.. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.1
3.1.2 Ductility . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.1.3 Volume Change Accommodation . . . . . . . . . . . . . . .
3.2
3.3 I
3.1.4 Durability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.4
3.1.5 Fire Resistance . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.1.6 Fabrication Simplicity . . . . . . . . . . . . . . . . . . . . . . .
3.1.7 Erection Simplicity . . . . . . . . . . . . . . . . . . . . . . . . . .
3.5
“3.: I
3.1.8 Appearance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.io
3.2 COnCeptSOf COMedOn DeSign . . . . . . . . . . . . . . . . . . . . . .
3.2.1 LoadTransfer Paths . . . . . . . . . . . . . . . . . . . . . . . . .
3.11
3.11
I
I
3.2.2 Failure Modes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.12
3.2.3 Stability and Equilibrium . . . . . . . . . . . . . . . . . . . . . 3.13
3.2.4 Stress Relief Measures.. . . . . . . . . . . . . . . . . . . . . . 3.14
3.2.5 fbqansion.Joints . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.16
3.3 Designing Load Transfer Mechanisms . . . . . . . . . . . . . . . . .
3.3.1 Bearing ..................................................................
3.3.2 Shear Strength
3.16
i.tz
.
I
I
3.3.3 Tensile Strength . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.26
3.3.4 Anchorage . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.26
3.3.5 Friction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.31

I
3.4 Load Transfer Devices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.32
3.4.1 Bolts and Threaded Connectors . . . . . . . . . . . . . . . . 3.32
3.4.2 Inserts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.37
3.4.3 Welded Studs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.44

I
3.4.4 Welding . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.47
3.4.5 Reinforcing Steel . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.53
3.4.6 Dowels . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.55

I
3.4.7 Post-Tensioning Steel . . . . . . . . . . . . . . . . . . . . . . . . 3.56
3.4.8 Pads and Other Bearing Devices . . . . . . . . . . . . . . . 3.57
3.4.9 Cast-In-Place Concrete . . . . . . . . . . . . . . . . . . . . . . 3.59

I
3.4.10 Grout . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.59
3.4.11 Epoxy Resins . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.61
3.5 Special Designs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.62
3.5.1 Reinforced Concrete Corbels . . . . . . . . . . . . . . . . . 3.62

I
3 S.2 Structural Steel Haunches . . . . . . . . . . . . . . . . . . . . 3.64
3.5.3 Dapped-endBeams . . . . . . . . . . . . . . . . . . . . . . . . . . 3.79
3.5.4 Hanger Connections . . . . . . . . . . . . . . . . . . . . . . . . . 3.78
References - Part 3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.84
I
TYPICAL CONNECTION DETAILS . . . . . . . . . . . . . . . . . . . . . . . . .
I
4. 4.1
4.1 Column to Foundation - CF . . . . . . . . . . . . . . . . . . . . . . . . . . 4.3
4.2 Column to Column - CC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . t.ti

I
4.3 Beam to Column - BC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.4 Slab to Beam - SB . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . i.Ei
4.5 Beam to Girder - BG . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

I
4.6 Beam to Beam - BB . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4:65
4.7 Slab to Slab - SS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.69

ii
I
I
I 4.8
4.9
Wall to Foundation - WP . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Slab to Wall-SW . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.75
4.82
4.10 Beam to Wall-BW . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.96
I 4.11 Wall to Wall-WW..................................4.10 1

References - Part4 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ..4.110


I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
...
111
I
ACKNOWLEDGEMENTS I
In addition to the listed authors, others on the staff of the Consulting Engineers I
Group who contributed significantly to this study were: K. R. Kowall, E. D. Losch,

A. J. Moore, G. E. Goettsche, and N. L. Scott. I


David A. Sheppard, chairman of the Prestressed Concrete Institute Seismic

Committee authored a draft report for that committee which was the basis for Section
I
2.5 of this report. D. P. Jenny, PC1 Technical Director was very helpful in digging up
I
I
resource material. Both he and G. D. Nasser, editor-in-chief of PC1 technical

publications contributed editorial assistance. J. R. Janney, Boris Bressler and others at

Wiss, Janney, Elstner and Associates added their experience from failure investigations

and earthquake research.


I
Finally, about 40 members of the Prestressed Concrete Institute, in particular, the
I
I
Committee on Connection Details, contributed literally hundreds of man-hours in review

and critique of Parts 3 and 4 of this report. With this help, the authors feel that this

I
document represents the latest state-of-the-art in precast concrete connection design

and detailing.

This material is based upon work supported by the National Science Foundation
I
I
under Grant No. PFR-7820900. Any opinions, findings, conclusions or recommendations

expressed in this publication are those of the authors and do not necessarily reflect the

I
views of the National Science Foundation or the Prestressed Concrete Institute.

I
I
This report is published and distributed by the Prestressed
Concrete Institute through agreement with The Consulting
Engineers Group, Inc., and with the forehowledge of National
Science Foundation. PC1 producer members, technical committees
and staff participated in the project as indicated in the
acknowledgements. However, the Prestressed Concrete Institute
cannot accept responsibility for incorrect engineering designs
I
I
resulting from errors or omissions in the report or misuse of
the matekal contained in the report.

I
iv
I
I
I
I
I
I
I
I P A R T 1

I INTRODUCTION

I
I
I
I
I
I
I
I
I
I
I
I
I 1. INTRODUCTION MD OVERVIEW

In a relatively short period of time, precast, prestressed concrete has become

I an important method of framing for structures. Virtually all types of structures are
being built with this material-industrial buildings, parking garages, commercial build-

I ings, multi-family housing, motels, schools, recreational buildings and bridges.


Advantages cited by the proponents of this method of construction stem

I
partly because the material is precast, partly because it is prestressed, and partly be-
cause it is concrete.

I
Precasting offers the opportunity of compressing construction time schedules.
For example, superstructure members can be cast and cured simultaneously with site
work and foundation construction. Erection of precast members takes less time compar-

I ed with cast-in-place concrete. Production of the products in a manufacturing facility


allows the use of accelerated curing techniques to further reduce time and produce

I higher strength concrete, improved quality control and better opportunities for standard-
ization.

I
Compared with ordinary reinforced concrete, the use of prestressing allows
longer spans with shallower depths, more controllable performance in terms of cracking

I
and deflection, and less material usage.
The principal advantages of concrete over other materials such as timber and
steel are fire resistance and durability.

I ized:
The disadvantages of precast, prestressed concrete can be similarly categor-

I Precasting of components allows less margin for error-requiring closer toler-


ances in casting. To fully realize the advantages, there must be adequate manufacturing

I
facilities within a reasonable hauling distance of the project.
Prestressing places concrete under constant compression, thus creep strains

I
are greater than in non-prestressed concrete. Eccentric prestressing, the most advanta-
geous structurally, causes cambers which must be considered in the design.
Compared with timber and steel, concrete members are heavier and bulkier,

I which can be a disadvantage in seismically active areas, or on sites with poor soils.
It should be noted that costs are not listed in any of the advantages or disad-

I vantages. Those experienced in construction recognize that no blanket statements can


be made regarding costs. Any construction material may be more or less cost competi-

I
I 1.1
I
tive for a given type of structure, depending on the design requirements, geographical
I
location, code restrictions, etc. It is to the advantage of the construction consumer to
have a choice of framing methods available, and to be assured that any method selected I
is designed and detailed in the most economical manner. It is axiomatic that the struc-
tures must be structurally able to resist loads that are likely to occur during the life of I
I
the structure.
1.1 Scope of this Study

I
This study represents the state-of-the-art, of precast concrete connection
technology. Special emphasis has been placed on the special considerations involved in

I
designing for earthquake resistance. However, the design methods of Part 3, and the
connection details shown in Part 4 apply to connections in general. It is intended that

I
those two parts serve as a general guide to improved design and construction of connec-
tions of precast and prestressed concrete. The special requirements for earthquakes and
other unusual loading such as hurricanes and tornadoes are contained in Part 2.
The team involved in this study has had extensive experience in the design and
construction of precast, prestressed concrete structures. Three consulting engineering I
I
firms were involved. These three firms have specialized in various aspects of the con-
struction method; design, detailing, field supervision and investigation of performance

I
and failures. The staff and members of the Prestressed Concrete Institute were also
deeply involved. Membership of PC1 includes companies and individuals who are actively
involved in the design, fabrication and construction of precast, prestressed concrete
buildings and bridges.
An extensive literature search and review was a major part of this study.
I
Bibliographies and references are given in each part. This report is a synthesis of the
pertinent information contained in the literature, supplemented by the hands-on exper- I
I
ience of the study team.
Interviews were conducted with people in the design and construction

I
community; building officials, engineers, architects, contractors and educators. Their
views as to the advantages, limitations and research needs were incorporated into the
study.

I
I
I
1.2
I
I
I 1.2 Development of the Precast, Prestressed Concrete Industry in

I 1.2.1
North America
History

I The basic concept of prestressed concrete was first patented in the United
States in 1886, but it was almost a half century later before the initial shortcomings

I were solved. P. H. Jackson’s original patent involved prestressing concrete by inserting


bolts through longitudinal holes in the concrete. The bolts were then tightened against

I plates bearing on the ends of the concrete member. Because shrinkage and creep can
cause about 0.1% shortening of the concrete, the initial stress in the steel diminished.

I
The bolts became loose and the concrete was no longer prestressed. Eugene Freyssinet, a
French engineer, demonstrated in the 1930’s that prestressing could be done successfully

I
if high strength steel were used. Thus if the concrete subsequently shrinks and creeps
0.1% of its length, the steel stress reduction (now called prestress losses) of 30,000 psi
could be accommodated, particularly if the initial steel stress was in the order of 150,000

I psi. Today most cold-drawn prestressing steels have an ultimate strength of at least
250,000 psi, and initial stresses of 175,000 to 190,000 psi are commonly used.

I After World War II, Europeans began to use prestressed concrete in their
rebuilding programs, principally because it afforded more efficient use of steel and

I
concrete. Most of the prestressed concrete used in Europe shortly after World War II was
post-tensioned rather than pre-tensioned, so it is not surprising that the first uses of

I
prestressed concrete in the United States employed post-tensioning. But shortly after
the first major project in the United States, the Walnut Lane Bridge in Philadelphia in
1951(l’l), Americans began to consider the advantages of plant-produced prestressed

I concrete.
In 1952 several prestressed concrete manufacturing plants, located mainly in

I Florida, Pennsylvania, Colorado and Washington, began producing pretensioned prestress-


ed concrete, not only for bridges but for buildings and other structures.

I
Several factors were responsible for the fact that plant-produced prestressed
concrete became popular in America before similar trends were apparent in Europe. Jn

I
America, good roads, a well developed trucking industry capable of hauling heavy bulky
loads, readily available mobile erection equipment, and high priced field construction
labor combined to make precast, prestressed concrete economical.

I From its modest start in the early 1950’s, the prestressed concrete industry
has grown steadily into a network of competing plants throughout the United States.

I
I 1.3
I
I
I
I
I
I
I
I
I
Fig. l-l - Estimates of dolkc sales of precast and prestresaed concrete
I
for the United States and Canada, 1950 through 1980
I
Fig. l-l shows the sales of precast and prestressed concrete in the United States and
Canada from 1951 through 1980(1*2). The growth rate between 1959 (sales of $100
million) and 1975 (sales $1.2 billion) averaged more than 16% annually. Since 1973 the
I
sales volume has fluctuated between about $1.1 and $1.8 billion annually, reflecting the
general volume of construction, excluding one- and two-story houses.
I
1.2.2 Development of standard products

In the early days of development of the precast, prestressed industry, many


I
types and shapes of precast concrete products were tried. Those which proved to be most
successful survived and became standardized. For bridges, two shapes of girders
I
emerged, l-shaped beams and box beams. ln the early 1960’s a joint committee of the
Prestressed Concrete Institute and the American Association of State Highway Officials I
I
developed standard cross section dimensions for both of these shapes as well as deck
slabs and piles(1’3). Standardized prestressed concrete bridge beams became available
throughout most of the United States, and many thousands of bridges have been built
with these sections. I
1.4
I
I
I Standardization of building products has developed more slowly. Because of
the relatively short shipping distance for precast concrete products, most owners of
I precasting plants consider their business to be local in nature, so standardization for its
own sake doesn’t offer much incentive. Furthermore, precasting plants sprang up almost
I simultaneously and began producing custom-designed elements. Because the design

I
profession had no standard sections, designers often developed sections that would best
solve a particular problem. In turn, the precaster purchased forms (often expensive steel
forms) to produce the product. That product then became a “standard” in the precaster’s

I area.
Some of the floor and roof slabs which were used on early projects included

I double tees, mono-wing tees, channels, single tees, quad-tees, solid flat slabs, and many
variations of these shapes. Through the years double tees proved to be the most econom-

I
ical for spans of 30 to 70 ft and have become standardized through industry efforts.
The earliest double tees were 4, 5, or 6 ft wide and were generally 12 to 18 in.
deep, with 14-in. deep by 4-ft wide double-tees gaining the widest usage. However, most

I double-tees made today are 8 ft wide and many are 9, 10, and 12 ft wide. Depths range
from about 12 in. to 40 in. The 24-in. deep by 8-ft wide section is the most widely used.

I Precast hollow-core slabs (made with non-prestressed bars) were made in


America prior to 1950. However the first machines for mass producing prestressed

I
hollow-core slabs were developed in Germany during the 1940’s. (The first such machine
was brought to the United States in 1954 in such poor condition that it was classified as

I
scrap on the trams-Atlantic crossing!) The growth of hollow-core slab production in the
United States was slow at first, probably because of the high capital investment required,
but during the 1960% and 70’s the use of hollow-core slabs increased sharply and now is

I the most used single product in the precast, prestressed concrete industry.
Hollow-core slabs were initially made in 4, 6, and S-in. depths, but today most

I units are 8 in. deep and many are 10 and 12 in. deep. Most hollow-core units have widths,
depths, and core shapes dictated by the manufacturing process. Most are machine-made

I
or are made in highly automated plants. They are generally marketed by the trade name
of the manufacturing equipment licenser, such as Spancrete, Flexicore, Span-Deck,

I
Spiro& Dy-Core, etc. Although the depths of hollow-core slabs have been standardized,
the widths are established by the manufacturing equipment. For example, Spancrete is
generally 40 in. wide, essentially the same as the width of the original German product,

I one meter. Flexicore is generally 24 in. wide, Spiro& Dy-Core, and Span-Deck are
generally 48 in. wide, but many Span-Deck and some Dynaspan slabs are 96 in. wide.

I
I 1.5
I
Precast concrete wall panels also constitute a large segment of the industry.
I
These panels can be loadbearing or non-loadbearing, and can be architectural and/or
structural. Manufacture of precast concrete wall panels predates the prestressing indus- I
try. Precast concrete wall panels were used on many notable structures prior to World
War II, and were referred to as “cast stone”, I’art stone”, and other similar terms. Many I
I
of these panels are still in use and serve as a testimony to the craftsmanship of the
builders and the durability of concrete panels.
Most of the early uses of precast wall panels were attempts to imitate natural
stone (limestone, sandstone, granite, or marble). Such panels were usually non-loadbear-
ing and were rather small in comparison to current practice. Connection details were
I
adaptations of those used for natural stone.
With the growth of the industry in America, the use of larger precast wall I
I
panels became economical. Wall panels of a variety of shapes, patterns, and textures
could be produced. These included custom-made architectural panels and %tandardtf wall

I
panels made of structural shapes, such as double-tees and hollow-core panels. Many of
these panels serve both architectural and structural functions. In recent years an in-
creasing number of precast concrete wall panels have been made as sandwich panels
incorporating an insulating material to help reduce heat losses.
In addition to floor and roof slabs and wall panels, a wide variety of other I
I
precast building elements are made. These include beams, columns, fascia panels,
column covers (for steel columns), elevator shafts, piles, and balconies. These items

I
have not been standardized by the industry, but most precasting firms generally make
only a small number of different sizes of each item.

I
Through the years the PC1 Committee on Standardization has been able to
develop certain standards, most notably those for double-tees. In addition, standards for
depths of hollow-core units and single tees have also been successfully developed. The
publication of the first edition of the PC1 Design Handbook(l’*) in 1971 helped to speed
up the standardization process, not only for cross sectional dimensions, but for design I
I
procedures as well.
1.2.3 Inhibitors to the Prowth of the industry
The growth of the precast prestressed concrete industry did not proceed
without problems. Indeed, the industry grew because the technical soundness of pre-
I
stressed concrete out-weighed the unsolved problems. A great deal of research on pre-
stressed concrete was conducted in the early 1950’s. Much of the research involved I
I
flexural behavior, deflection and camber prediction, and bond and anchorage stresses.

1.6
I
I
I Later researchers studied end-block stresses, behavior under repeated loads, shear
strength and prestress losses. In the early 1960’s, systems studies were begun of fire
I resistance, long term durability, restrained shortening stresses, and certain connection
details. Research on some of these topics, most notably connections is still incomplete.

I One of the problems faced and solved by the industry in the 1950’s was the
lack of building code provisions. Prior to publication of the 1963 AC1 Building Code

I Requirements for Reinforced Concrete, which included provisions for prestressed


concrete, an ACI-ASCE Joint Committee published a Recommended Practice in

I
1958(1’5), and in 1959, PC1 published building code provisions (1.6). However, many
building officials were reluctant to approve the use of prestressed concrete until the

I
local building codes adopted the 1963 AC1 Code(1’71.
Another problem faced by the young industry was that of fire resistance.
Because prestressed concrete is made with cold-drawn prestressing steel which is affect-

I ed more by high temperatures than other steels used in construction, questions about the
fire resistance arose. In 1957, PC1 organized a committee on fire resistance, and in 1958

I
the first full-scale standard fire tests were conducted at Underwriters Laboratories.
Through the years, the industry has sponsored both standard tests and comprehensive

I
research programs, to the point that more data exists on the behavior of prestressed
concrete exposed to fire than most other construction materials (1.8). In recent years, a
methodology for calculating fire endurance of prestressed concrete has been developed.
I The PC1 publication, “Design for Fire Resistance of Precast Prestressed Concrete t&9 ,
is accepted by most building officials as an alternate method to fire testing for deter-

I mining the fire endurance.


Much of the economy of precast, prestressed concrete structural framing lies

I
in its simplicity. It is best usad in simple span, pin-ended beams and deck members. The
absence of continuity and redundancy has caused some designers to question stability

I
under high lateral loads.
Precast, prestressed concrete has been used in many areas throughout the
world which are subjected to high winds (hurricanes, typhoons, and tornadoes) and to
I earthquakes. Many thousands of buildings, bridges, stadiums and other structures have
been built of precast, prestressed concrete in hurricane-prone regions and in seismically

I active areas of the western United States, Alaska, Japan, New Zealand, Indonesia and
along the Mediterranean Coast, as well as in western South America. Many of the struc-

I
tures have successfully withstood hurricane forces and earthquake motions. Some
precast concrete structures have been damaged by earthquakes. Overall behavior of

I
precast concrete structures in earthquakes has been about the same as that for steel or

I 1.7
I
cast-in-place concrete structures. However, the number of precast concrete buildings I
that have been involved in serious earthquakes is relatively small, so it may be premature
to make any general conclusions. This aspect of precast, prestressed concrete construc-
tion is one of the primary purposes of this study. It is discussed in detail in Part 2.
I
Probably the most persistent problem facing the industry has been that of
adequate connections to withstand the variety of stresses and movements imposed by
I
gravity loads, repeated loads, creep and shrinkage forces, temperature changes and
intermittent lateral forces of wind and earthquake. Early attempts at developing ade- I
I
quate connections resulted in either go-no-go details or complex empirical expressions
that did not relate directly to imposed shears, moments, and direct forces. The relative-

I
ly recent concept of ‘shear-friction” has been of considerable help in developing rational
design methods to replace empirical relationships. But the problem of providing econom-
ical connections that will transmit the required forces without unduly restraining the
volume change movements which occur still persists. The other primary thrust of this
report is to synthesize available information on connections design and detailing. These
I
aspects are covered in Parts 3 and 4.
Despite these shortcomings, the prestressed concrete industry has had an I
I
impressive growth record. Failures which have occurred have been well publicized and
the industry has worked to prevent similar Occurrences by disseminating information to

I
designers.

I
I
I
I
I
I
I
1.8
I
I
I .
RJWERENCES - PART 1

I 1.1 Zollman, Charles C., ~~Reflections on the Beginnings of Prestressed Concrete in


America-Part 1, Magnels Impact on the Advent of Prestressed Concrete”, Journal
I of the Prestressed Concrete Institute, v. 23, no. 3, May-June, 1978.
1.2 (Annual report, to PC1 Membership, 1980).
I 1.3 Joint Committee - American Association of State Highway Officials Committee
on Bridges and Structures and Prestressed Concrete Institute (Available from PC1

I - Chicago)
(1) “Standard Prestressed Concrete Beams for Highway Bridge Spans 30 ft to

I
140 ft”.
(2) “Standard Presstressed Concrete Box Beams for Highway Bridge Spans to
103 Feet”.

I
(3) “Standard Prestressed Concrete Slabs for Highway Bridge Spans up to 55
Feet”.
(4) ‘Standard Prestressed Concrete Piles lo”, 12”, 14”, 16”, 18” 20”, 22”, and
24”“.

I
“Prestressed Concrete Cylinder Piles 36” - 48” - 54”“.
I:; “General Notes - Prestressed Concrete Piles”.

I
1.4 “PC1 Design Handbook - Precast and Prestressed Concrete”, First Edition, 1971,
Prestressed Concrete Institute, Chicago.
1.5 ACI-ASCE Joint Committee 323, “Tentative Recommendations for Prestressed
I 1.6
Concrete”, AC1 Journal, v.29, no. 7, Jan. 1958.
“PC1 Standard Building Code for Prestressed Concrete (Tentative)“, Std. llO-59T,

I
Prestressed Concrete Institute, Chicago.
1.7 “Building Code Requirements for Reinforced Concrete (AC1 318-63Y’, June, 1963,

I
American Concrete Institute, Detroit.
1.8 “Fire Resistance DirectoryI’, Jan, 1980, Underwriters Laboratories, Northbrook, IL.

I
1.9 Gustaferro, A. H., and Martin, L. D., “PC1 Design for Fire Resistance of Precast,
Prestressed Concrete”, 1977, Prestressed Concrete Institute, Chicago, IL

I
I
I
I
I 1.9
.
I
I
I
I
I
I
I
I
P A R T 2

I
KARTHQUAKIS AND OTHER EXTREME LOADING

I
I
I
I
I
I
I
I
I
I
I
I
I 2. EARTHQUAKE AND OTHER EXTREME LOADING

2.1 Introduction
I An earthquake is one of nature’s most unpredictable and devasting

I
forces. The energy release causes forces which often are greater than the structural
resistance of natural and man-made structures. Loss of life, structural collapse, or

I
excessive structural damage often occur unnecessarily in major earthquakes. The pre-
vention of these unfortunate events can only occur if proper design, detailing and quality

I
control are incorporated into the structural system. The present state of knowledge in
earthquake-resistant design has produced economical structures which have performed
safely in major earthquakes, but further research and advancement is still required. This

I is especially true when precast and prestressed concrete elements are used in earth-
quake-resistant structural systems.

I The major emphasis in earthquake resistant design has centered around the
use of steal and cast-in-place concrete as the major structural materials. Ductility,

I
energy dissipation, and the ease of producing monolithic redundant structures has made
these materials the forerunners in earthquake-resistant research and construction.

I
In the United States, the popular acceptance of prestressed and precast
concrete as a seismic lateral load resisting system has generally been limited to low rise
panel structures. Although this trend is changing, the U. S. is still conservative compar-

I ed with other seismic regions of the world such as Japan, Cuba, and Eastern Europe,
where multi-story construction of large precast panel buildings is widespread (2.33).

I The reluctance to use precast concrete in highly seismic regions of the


U. S. is partially because of the lack of research and performance data. Damping values,

I
post-elastic behavior, and cyclic strength of connections are dark areas which lack
information. As a result, design provisions and code requirements are often ambiguous,

I
vague, or non-existent regarding precast construction. Consequently, building officials
and design engineers often interpret code requirements differently, which results in
costly delays or total rejection of the precast system. Knowing this, owners and design-

I ers tend to prefer cast-in-place concrete and steel in seismic areas, since design con-
cepts and code requirements are well established and clearly documented.

I
The major building codes such as AC1 318-77(2’25), or the Uniform Building
Code(2.15) , have special provisions for the seismic design of reinforced concrete struc-

I 2.1

I
I
tures. They do not, however, have corresponding provisions for prestressed or precast I
concrete structures. The present design codes generally require precast systems to
conform to the requirements of reinforced concrete even though these requirements I
were developed for cast-in-place concrete and in many instances do not apply to precast
systems. I
I
The remainder of this section discusses some basic engineering principles
associated with the code requirements for reinforced concrete. Emphasis is placed on
the requirements for connections and the validity of applying reinforced concrete re-
quirements to precast and prestressed systems. The section is intended to provide a
better understanding of connection design for the engineer with limited experience in
I
I
seismic or precast-prestress design. The section also describes what little data there is
available on precast seismic behavior, and areas in need of future research.
2.2 Nature of Earthquake Forces on Buildings
A map showing the location of major seismic activity is shown in Fig. 2-l.
I
Highest seismic regions include the coasts of the Pacific Ocean from Chile through
Central America, along the coasts of the United States, through the coastal islands of I
I
British Columbia, southern Alaska, the Aleutians, Japan, the Philippines, New Guinea,
and New Zealand. Other major areas of seismic activity include central and eastern

I
United States, the Mediterranean coast, central Asia, and most of Indonesia. These areas
and other areas with previous seismic activity are major areas of concern for earthquake
resistant structural design.
The most widely accepted explanation of the origin of earthquakes is the
“elastic rebound theory” proposed by H. P. Reid in 1906. Reid claimed the occurrence of
I
I
an earthquake is due to a sudden shear fracturing, or faulting along planes of weakness in
crustal rock. The exact forces which produce the internal fracturing strains is not fully

I
known, but viscous movement of crustal plates is believed to be a major source.
When crustal rock fractures, energy is released in the form of a dynamic

I
wave. The vibratory ground action produced by the passing of this wave is what consti-
tutes an earthquake. When fracturing occurs along a number of different planes the
dynamic waves interact in a complex random motion. The magnitude of ground oscilla-
tion associated with this motion will depend on the ground material, the medium through
which the wave traveled, and the distance the wave traveled. Thus, the ground motion I
for a given earthquake will vary with the structural site.
The ground motions produced during an earthquake are recorded in the I
I
form of ground accelerations. Accelerations in both the horizontal and vertical direc-

2.2
I
I .-. .,.

-
i -4% t

%- 2-l - Map shone global seismicity for the year 1966 (Ref. 2.2)
0.3r I,
0.2
01
:: 0
-01
-0.2
-0.3 Ground Acceleration, y

Fig. 2-2 - Recorded Accelerogram for the El Centro, California earthquake


of May 18,1940, N-S component (Ref. 2.3)
I
tions have been recorded during major earthquakes in the last few decades. Fig. 2-2
shows the recorded accelerogram for the El Centro, California earthquake of May 18, - I
I
1940. This earthquake is one of the most intense long duration strong-motion earth-
quakes that has been recorded, and the accelerations shown are often used for structural

I
design.
Generally, the magnitude of horizontal ground acceleration in a major
earthquake will range from 0.2 to 0.5 g, while vertical ground accelerations hardly ever
exceed 0.1 g.
2.3 Predicting Structural Behavior
I
Two procedures are used to design structures to resist earthquakes. High I
I
rise or complex shaped buildings may be analyzed dynamically, using a predicted or
recorded ground motion. Most common, however, is to apply equivalent static loads
which are prescribed by building codes.
2.3.1 Dynamic Analysis Procedures
I
I
Dynamic analysis of a structure can be very complex and a detailed treat-
ment of the subject is beyond the scope of this report. There is even some disagreement
among design engineers as to the value of complex dynamic analyses, given the unpre-
dictable nature of earthquake movements and the necessity of making what may be
rather broad assumptions as to the response of foundation materials, the capacity for
I
damping, the post-elastic behavior of the structural components, etc.
Nevertheless, all modern methods of earthquake analysis are based on at I
I
least approximations of dynamic forces, so a brief overview of dynamic analysis proce-
dures is useful in understanding the various design techniques, including equivalent static
loads.

2.4
I
I
I The various dynamic analysis procedures determine the structural response

I of a system by solving the equations of motion for an acceleration applied at the base of
the structure. The design acceleration may be determined from an accelerogram record
of an actual earthquake, Fig. 2-2, or from a simulated motion generated from a geolog-
I ical study of a given site.

I
Three main methods are currently used for dynamic analysis. They are:
1. direct step-by-step integration of the equations of motion

I
2. normal mode analysis
3. response spectrum method

I
2.3.1.1 Direct Integration
Direct integration of the equations of motion is the most sophisticated

I
analysis method available for earthquake design. It is a step-by-step procedure which
determines forces and displacements by summing the response of the structure during
very short time increments. The structure is assumed to be linearly elastic during each
I increment of time, but between increments the properties of the structure are modified
in accordance with the current deformation condition.

I interval(2’4).
The following steps summarize the analysis procedure during each time

I
1. The stiffness of the structure for the time interval is evaluated,
based on the state of displacement existing at the beginning of the

I
interval.
2. Changes in displacements are computed, assuming the accelerations
to vary linearly during the interval.

I 3. These incremental displacements are added to the displacement


state at the beginning of the interval to obtain the displacements at

I 4.
the end of the interval.
Stresses are computed from the total displacements, taking into

I
account non-linear material properties.
The accuracy of the direct integration procedure is dependent on the

I
analysis method, the ground motion, and the modeling of the non-linear properties. For
precast concrete, little information is available on the linear or non-linear dynamic
behavior of connections. Thus non-linear analysis by direct integration is not as common

I for precast structures as it is for steel or monolithic concrete structures. ‘lhe majority
of non-linear analyses that have been performed for precast concrete involve the behav-

I ior of large panelized buildings(2.5, 2~3, 2*7)

I 2.5
I
I
I
I
I
First (fundamental)mode Second mode Third mode

Fig. 2-3 - Typical shapes of the fit three modes of vibration

I
I
2.3.1.2 Normal Mode Analysis
Normal mode analysis is more limited than direct integration analysis,

I
since non-linear behavior cannot be taken into account. Like direct integration, normal
mode analysis applies earthquake accelerograms to the structure and uses a stress history
to determine the structural response. However, unlike direct integration, the normal
mode analysis procedure uncouples the different modes of vibration for the structure.
The accuracy of the analysis is then affected by the number of modes which are superim-
I
posed.
A structure will generally have as many modes of vibration as it has signif- I
I
icant degrees of freedom. For the elastic response of multi-story buildings the first or
fundamental mode contributes about 80% of the total structural response, while the

I
second and third modes contribute about 15’%6(2*1). Thus sufficient accuracy can be
obtained with the superposition of the first few modes. Fig. 2-3 shows the first three
natural modes of vibration for a multi-story structure.
2.3.1.3 Response Spectrum I
I
The simplest of the dynamic analysis procedures uses a predetermined
a “response spectrum”. A response spectrum relates the maximum responses (acceleration,
velocity or displacement) of a structure to its natural frequency of vibration or period.
(Note: the velocities and accelerations shown in a response spectrum are not the actual
values, but values chosen for convenience in analysis. They are usually called ‘pseudo-
I
velocity” or “spectral velocity” and “pseudoacceleration” or “spectral acceleration”.)
Fig. 2-4 shows combined response spectra developed from the ground motion recorded
I
during the El Centro Earthquake. Fig. 2-S is an idealized version of the acceleration
response spectra for the same earthquake, which is more useful for design. I
I
The values shown in a response spectrum represent the maximum response

2.6
I
I
I
I
I
I
I
I
I
I
2-4 - Response spectra for El Centro Earthquake of May 19, 1940, N-S co

I
I
I
I
I
I
I Period ol vibration, T ( oec)

I pig. 2-5 - Av‘wage acceleration response spectra, 1940 El Centro intensity

I
I 2.7
I
I
Mass

I
/////////i-t_.
t
I
‘x 1 Damper X- Ground displacement
I
Original
Position
:
7
A
D: Mass displacement I
I
:

‘7r///////////////////////////////////////// I

Fig. 2-6 - !Sngle degree of freedom structure I


of a single degree of freedom structure, such as the frame modeled in Fig. 2-6, when the
base is subjected to the assumed earthquake motion. The spectra may be used in analyz- I
I
ing most multi-degree of freedom systems by treating each response mode separately.
As the structure passes from one mode to the next, the natural frequency (period), stiff-

I
ness and damping characteristics change. The total reponse is obtained by applying a
“modal participation factor” and summing the responses, similar to the method used in
the nor ma1 mode analysis.
Pseudovelocity, V, and pseudoacceleration, A, are related to the displace-
ment D, of the system by the following equations:
I
A =W2D (2-la)
(2-lb)
I
I
V~=wD
Where w i,s the circular frequency of vibration for the system and is defin-

Q-2)
I
I
where K = stiffness of the system
m. = mass of the system

I
The natural frequency, f, and period, T, of the system are given by the
relationship:

To illustrate the use of the response spectrum technique, assume the frame
I
in Fig. 2-6 is to be analyzed for the motion produced by the El Centro Earthquake. If the
stiffness of the frame is 25 kip/in. and the mass is 0.6 kip-sec2/in., then the circular I
I
frequency and period of the frame are:

2.8
I
I
I
w = E= r& = 6 . 4 5 radlsec
I
f = k = !y = 1.03 cps
I
I
T = + = 0.97

Lf 5% critical damping is assumed, the value of pseudoacceleration given in Fig. 2-5 is

I 9 ft/sec2.

I The maximum displacement is:

I
D = if = -$~f = 0.22 ft = 2.6 in.

I
The maximum base shear can be computed in two ways:

P max = k~ = 25c2.6) = 65 kips

I P max = mu = 0.6(9)(12) = 64.6 kips

I The response spectra for the El Centro earthquake and response spectra for

I
other dynamic motions usually have the following characteristics (2.3).
1. For very low frequencies the maximum response displacement, D, is

I
virtually constant and is equal to the maximum ground displacement.
2. For very high frequencies the maximum pseudoacceleration, A, of
the mass is virtually constant and is practically equal to the maxi-

I 3.
mum acceleration of the ground.
For intermediate frequencies, the maximum response displacements,

I velocities and accelerations are all amplified over the ground motion
maxima. For damping values in the range of 5 to 10 percent, the

I
amplification factors are approximately 1.0, 1.5 and 2.0 for dis-
placement, velocity and acceleration.

I
Fig. 2-5 shows that the largest accelerations, or forces for elastic response,
will occur in structures with natural periods of vibration around 0.5 seconds. Periods
greater than this produce much lower accelerations, white shorter periods tend to have

I
I
2.9
I
accelerations equal to or slightly less than the maximum.
I
The majority of low to medium rise precast concrete structures have
periods less than 0.3 secondsf2*10). This places them to the left and near the magnitude I
of maximum acceleration. If stiffness degradation occurs, the period of the structure
will increase but the accelerations will also increase. ‘Ihis is unlike structures with I
I
periods greater than 0.5 seconds where an increase in the structural period will decrease
the induced acceleration.

I
2.3.1.4 Damping
Figs. 2-4 and 2-5 show that damping has a major influence in determining

I
the structural response of a system. Reference 2.11 defines damping as “a measure of
the ability of a member or a structure to absorb the energy generated by the application

I
of an external repetitive type loading. ‘Ihis energy absorption reduces the oscillations
resulting from the loading and is expressed as a percentage of critical damping, which is
the minimum viscous damping that will allow a displaced system to return to its initial
position without oscillation”.
The exact value of damping for most structures cannot be determined since I
I
many structural and non-structural effects contribute to damping. Friction between
members, cracking of walls, yielding of members and connections, and soil conditions all

I
contribute to damping in the structure. Damping will also vary with the structural form
and the nature of vibration. Large amplitude post elastic vibration is more heavily
damped than small amplitude vibration (2.12) .
Damping values used for design are obtained through extrapolation of test
data for individual components and structures. Typical design values for various con-
I
struction types are given in Table 2-1f2.12). Generally, buildings with heavy shearwalls,
heavy cladding, or partitions, have greater damping values than lightly clad skeletal I
I
structures<2*12).
The high damping values given for concrete construction may not apply to

I
prestressed concrete since these values are dependent on the member behavior and the
connections between the members. Generally, the high damping values associated with

I
concrete are a result of tensile cracking in the member. For prestressed members,
cracking will not occur at working stress level, therefore the initial internal damping will
be lower than that produced by a reinforced concrete member. This can be seen from
the test results given in Table 2-2(2*27). Damping will increase as the member cracks,
but the connections between members must still withstand the structural deformations I
I
associated with the initial low damping values.

2.10
I
I
I ‘lhble 2-l Typical Damping Ratios

Type of Construction Damping


I (percent of critical)

Steel frame, welded, with all walls


I of flexible construction 2

I
Steel frame, with concrete shear walls 7

Concrete frame with all walls of flexible


construction 5

I Concrete frame, with concrete or masonry


shear walls 10

I Concrete shear wall building 10

I
I
I
Type of Member Damping
(percent of critical)

Frestressed Without tension cracks

I
0.5 to .75
concrete
beam With tension cracks not
visible to naked eye 1.0 to 2.00

I Reinforced
concrete
Without tension cracks 1.0

I
beams With first tension crack 2.0

2.3.1.5 Non-Linear Response Spectrum

I Response spectra can be developed for well defined non-linear behavior.


Fig. 2-7 shows such a spectra presented by Newmark for an elasto-plastic system with a

I damping value of 2% of criticaL


Ihe validity of modeling precast concrete as an elasto-plastic system is

I questionable. However, the general characteristics of the systems are beneficial in


understanding non-linear structural behavior. ‘lhe following generalizations were noted

I by Newmark(2’3).
1. For low frequency systems, the total displacement for the inelastic

I
system is the same as for an elastic system having the same frequency.

I 2.11
I
I
I
I
I
I
I
I
I
Fig. 2-7 - Deformation spectra for elaeto-pleetic systems with 2%
eciticel damping subjected to IQ Centro Earthquake (Ref. 2.3)

2. For intermediate frequency systems, the total energy absorbed is the


same for the inelastic system as for an elastic system having the same
I
3.
frequency.
For high frequency systems, the base shear is the same for the inelastic
I
2.3.2
system as for an elastic system having the same frequency.
Seismic Building Design Codes
I
The most common method of obtaining forces and designing precast struc-
tures for earthquake loads is with the “equivalent static force method” given by most
I
building codes. ‘Ihe method allows an engineer to approximate earthquake forces with
coefficients and formulas based on the physical properties of the structure rather than
I
with a complex dynamic analysis.
The current seismic design codes in the United States were developed largely I
I
through the work of the Structural Engineers Association of California (SEAOC). Their
report, “Recommended Lateral Force Requirements and Commentary r,(2.13) , is the major

I
source for the seismic design requirements of the Uniform Building Code, the most
commonly used code for earthquake design. It is not surprising that the SEAOC is a
major contributor to seismic design codes, since the majority of seismic activity and
earthquake resistant design in the U. S. is centered in California. The codes and prac-
I
I
2.12
I
I tices in California are highly regarded in the area of earthquake resistance design.
TherefOre, a brief review of early California design codes and the developments of these

I codes will be beneficial in understanding the present code system.


2.3.2.1 Early Developments
I One of the first major earthquakes in the United States to cause major loss of

I
life and damage occurred in San Francisco in 1906. Statistics regarding the earthquake
are questionable, but it is believed that 700 to 800 people died and over 400 million

I
dollars in damage occurred from the earthquake and the fires after the
earthquaket2*14). Surprisingly, the earthquake and the damage which ensued did not
produce any changes in the building codes. The city was rebuilt to a design code that

I treated wind and earthquake alike as a horizontal pressure of 30 psf(2*40).


After the 1925 Santa Barbara, California earthquake which caused 12 deaths

I and about 6 million dollars in structural damage (2.14) , structural engineers in California
began to give serious thought to earthquake resistant design. Research was started in

I
the United States and design procedures began to appear in engineering journals. How-
ever, there were still no mandatory earthquake design requirements in any building

I
codes. It was not until the Long Beach, California earthquake in 1933 that earthquake
design requirements appeared in building codes in the United States(2’g)
The first building code provisions in the U. S. regarding earthquake resistant

I design were similar to the provisions of Japanese codes which had been in existence for
almost ten years. The first codes specified a single seismic coefficient with the earth-

I
quake forces calculated by the following formula:
F=CW (2-J)

I
Where F is the earthquake force, C is the seismic coefficient and W is the total dead load
of the structure plus one half of the design vertical live load.

I
In the 1933 codes, C was a constant equal to 0.08 for ordinary buildings and
0.10 for school buildings. Factors such as construction type, building height, or structur-
al shape, did not affect the value of C. However, certain individual parts of buildings

I such as parapet walls, chimneys, tanks, and exterior ornamentation were designed for
higher earthquake forces since experience had shown these items to be susceptible to

I earthquake damage (2.9).


In 1943 the code for the City of Los Angeles revised the value of C to ac-

I
count for the fact that the period of vibration for a building will increase with the num-
ber of stories, and smaller accelerations are obtained with large periods of vibration.

I 2.13

I
I
This can be seen from the acceleration spectra in Fig. 2-5. The new value of C was given
I
as:
I
60
’ = IN + 4.51100 (2-5)
I
where N is the number of stories above the one under consideration.
The value of C was revised again in 1957 after the City of Los Angeles re-
moved the 13 story height limitation which had been imposed on previous building con-
I
struction. The new equation for C was:
I
c = w+ 0.9”;?- S)J 100 (2-6)

where S is the total number of stories in the building. For buildings less than 13 stories,
I
the value of S was taken as 13 and the value of C remained unchanged from the previous
code value.
I
I
The new equation for C produced design forces which gave a better approxi-
mation to the forces determined by a dynamic analysis for buildings greater than 13

I
stories. It also was the first equation to approximate the triangular load distribution
predicted by dynamic analysts.
The first seismic committee of the Structural Engineers Association of Cali-
fornia was formed in 1957. The intent of the committee was to develop a seismic code
which would be in close agreement with the essential features of dynamic theory, and to
I
develop a uniform code which could replace the many different codes that existed at the
timet2sg) . I
I
The code developed by SEAOC was the first to incorporate the concept of
ductility into the design process. It was also the first to impose different requirements

I
for different construction types.
The base shear derived by the committee was given by the formula:

V=KCW (2-7)
I
In which V is the base shear, K is a ductility factor, W is the total weight of the building
above the first floor and C is the base shear coefficient defined by the equation: I
(2-E) I
I
2.14
I
I
I In this equation, T is the fundamental period of vibraton of the building. The committee
realized the calculation of the period of vibration for a building was complicated and
I time consuming. Therefore, they developed the following simple formula for calculating
T:
I T = 0.05H
6
(2-9)

I Here H is the height of the building in feet and D is the dimension of the building, in
feet, in the direction parallel to the applied forces. The code stipulated that T need not

I be less than 0.10 for the purpose of calculating C.


The formula for T did not have a high degree of accuracy since it represented

I average values obtained by measuring periods of vibration for several hundred buildings.
However, it was felt that the errors in the period of vibration would not seriously affect

I
the calculated base shear value.
The ductility factor, K, accounted for the difference in resistance to seismic

I
loads for different types of construction. The values of K were based on the opinions of
the committee members and on the performance of each system in previous
earthquakes. Types of construction which had performed well in earthquakes were

I assigned low K values, while structures with poor performance records were given high K
values. The four K values given by the committee are listed below (2.9) .

I
K= 1.33 was established for the Wax Type” building structures which do not have
a complete space frame capable of resisting all vertical loads. Vertical and

I
lateral loads are carried by bearing walls and shear walls.
K= 1.00 was established for a building having a complete “vertical load space
frame” designed to carry all vertical loads. Shear walls or other bracing
I K=
systems are designed to resist the total lateral force.
0.80 was established for buildings with a complete horizontal bracing system

I capable of resisting all lateral forces, plus a ductile moment resisting space
frame which independently is capable of resisting a minimum of 25 percent of

I
the lateral force.
K= 0.67 was established for buildings in which the total lateral force is resisted

I
by a moment resisting frame which has the necessary ductility.
The code also developed the following formula for determining the shear

I
force at each story level:

(2-10)

I
I 2.15
I
where:
F, = lateral force at level x. I
v zz base shear
WX
= weight at level x I
hx = height in feet above the base to level x
cwtl = summation of the products of all wx h,‘s for the building I
For determining overturning moments the committee recommended the
forces at each floor level be reduced by a factor J. This accounted for the reduced base
moments produced during high modes of vibration when the floor shear forces are not
I
acting in the same direction. The factor was later dropped when damage from the 1967
Caracus, Venezuela earthquake showed results obtained with a J factor were unconserva- I
I
tive.
The code developed by the 1957 SEAOC committee was first published in

I
1959. Since that time the code has been changed to account for lessons learned in later
earthquakes. However, the major principles and the general approach to earthquake

I
design developed by the committee is basically the same in the current codes.
2.3.2.2 Present Code Requirements

I
The earthquake provisions of the 1979 Uniform Building Code represent the
current equivalent static design requirements. The base shear formula given by the UBC

I
has been revised to the following formula:

V = ZIKCSW (2-11)

Z is a zone factor coefficient which accounts for the probability of an earth-


I
quake occurring in a given region. The coefficient is based on previous seismic activity
in a region, with regions of highest probable activity, or major fault systems, having a Z I
I
value of 1. Figure 2-8 shows the different zones in the continental United States. The
values of Z are 0, 3/16, 3/8, 3/4 and 1 for zones 0 - 4 respectively.

I
The term I is an occupancy importance factor which establishes higher seis-
mic forces for structures essential to public safety during and after an earthquake.
Essential facilities have an I value of 1.5 and include hospitals, medical buildings with
surgical facilities or emergency treatment areas, fire and police stations, disaster oper-
ation facilities and communication stations. Any building with a primary occupancy of
I
I
more than 300 people in one room has an I value of 1.25, while other buildings have an I
value of 1.0.

I
K, the ductility coefficient, is basically the same as the values defined by the

2.16
I
I
I
I
I
I
I-
I
I
I
I
w. or n * L c

pis. 2-8 - Seismic zone map of the United States (Ref. 2.15)

I 1957 SEAOC committee.

I
The value of C is currently given by the formula:

I
c = & so.12 (2-12)

For buildings with a ductile moment-resisting space frame capable of resisting 100 per-

I cent of the lateral load, and without being enclosed or adjoined by more rigid elements
tending to prevent the frame from resisting lateral forces, the period can be found by the

I
formula:
T = O.lON (2-13)

I
Other buildings may use the formula given by the original committee:
o.05Hn
T=- (2-14)

I
6

S is a coefficient which accounts for the site-structure resonance. S is de-

I termined by the following formulas:

1~ 2.17
I
For T/Ts = 1.0 or less: I
(2-15) I
For T/Ts greater than 1.0: I
S = 2.3 + 0.6 6 -0.3 & 2
S [IS I
If Ts, the characteristic site period, cannot be properly established, the value of S should
be taken as 1.5. The code also specifies S should always be greater than 1.0, and the I
I
product CS need not exceed 0.14.
The UBC also gives the following formula to determine the forces on parts or

I
portions of structures, non-structural components and their anchorage to the main struc-
tural system:

Fp = ZICpWp (2-17)
I
Fp is the force on the component, Wp is the weight of the component and Cp
is defined in Table 23-J of the UBC. For most precast concrete components and their I
I
connections, Cp = 0.3.
The body of a precast connection for non-bearing non-shear wall panels is

I
required to resist 4/3 Fp. Fasteners attaching the COMEXtOr to the panel or the StrUC-
ture such as bolts, inserts, welds, or dowels are required to resist 4Fp.
Connections and panel joints should also allow for a relative movement be-
tween stories of the greater of l/2 inch, two times the story drift caused by wind,
Or (y) times the calculated story displacement caused by seismic forces.
I
Braced frames in buildings located in Seismic Zones No. 3 and No. 4 and
buildings located in Seismic Zone No. 2 with an importance factor, I, greater than 1.0, I
I
are required to design their members and connections for 1.25 times the calculated
earthquake force.

I
2.3.2.3 Future Code Provisions
In 1878 the Applied Technology Council, in cooperation with the National

I
Bureau of Standards and the National Science Foundation published a comprehensive
report on the current state of knowledge in the fields of engineering seismology and

I
seismic design for construction of buildings. The report, commonly called ATC-3(2’16),

2.18
I
I
I was developed for the improvement of current building practices. ATC-3 has not been
formally adopted and present debate over some of the provisions will probably produce
I revisions in the report. However, if the majority of the existing report is adopted into
the major building codes, earthquake design requirements will change for a large portion
I of the country.
Many of the changes which ATC-3 will produce stem from the revision of the

I different seismic zones throughout the country. Seven zones are listed in ATC-3 instead
of the four found in UBC, and many areas are placed in new zones which have more

I stringent earthquake requirements than the present UBC requirements. This is especially
true for many midwestern and eastern areas where earthquake requirements currently

I
have an insignificant effect on the structural design. However, under ATC-3, earthquake
resistance would be a major factor in design, significantly affecting cost, material selec-

I
tion and construction methods(2’17).
.
Since the provisions of ATC-3 are still tentative, a detailed discussion of the
report will not be given. Relevant provisions concerning connection design will be dis-

I cussed in later sections. The significant factor is that the present UBC earthquake
provisions may be revised to incorporate higher design forces along with stricter ductility

I and continuity requirements.


2.4 Design Philosophy

I The major building codes consider the following criteria as a basis for struc-
tural performance:

I 1) Resist minor earthquakes without damage. A structure would normally


resist such frequent but minor shocks within its elastic range of stress.

I 2) Resist moderate earthquakes without significant structural damage, but


with some non-structural damage.

I
3) Resist major catastrophic earthquakes without collapse. Essential
facilities are also expected to remain operational during and after a
major earthquake.

I The previous sections have shown that once the physical parameters of a
structure are known, approximate earthquake forces can be generated through a dynamic

I analysis or simplified code equations. The equivalent static forces prescribed in the
major building codes were developed with the assumption that inelastic structural re-

I
sponse will occur during a major earthquake. Structures which are designed for elastic
behavior with these static loads will be overstressed during a major earthquake. Suffi-
cient ductility is required to accommodate the deformations associated with inelastic
I
I 2.19
I
behavior without jeopardizing the integrity or stability of the structural system. Using a
I
dynamic analysis, the engineer has the option of assuming elastic or inelastic structual
behavior. For elastic behavior, the design forces approximate the maximum I
anticipated. Little damage will occur if strength degradation is prevented and structural
deformations are controlled. If the structure responds inelastically, lower design forces I
I
are used, resulting in smaller members, less weight and a more economical structure.
The energy principles discussed in the next section will describe ductility
requirements and inelastic behavior of simple structural systems.
2.4.1 Energy Principles I
I
The easiest way to Illustrate non-linear behavior in a structure ls through the
oscillations of a simple cantilever as shown in Fig. 2-9. Such an oscillator, responding
elastically, will have a load deflection relationship as shown in Fig. 2-9a.
The potential energy stored in the elastic system at the maximum deflection
point b, is equal to the area a-b-c under the curve. As the mass returns to its initial
I
position, the potential energy is converted into kinetic energy.
If the stress in the oscillator exceeds the yield strength of the material, a I
I
plastic hinge forms and the inelastic load-deflection will be as shown in Fig. 2-9b. At the
maximum response of this system, point e, the potential energy stored in the system Is

I
equal to the area a-d-e-f. When the mass returns to the zero position the area a-d-e-g
represents the energy dissipated by the plastic hinge. ‘lhe area e-f-g is the potential
energy converted into kinetic energy(2.18).
The energy dissipation capacity of the plastic hinge, or of any structural
component, dictates how the structure will respond to seismic ground motion. If the
I
energy dissipation is large, the vibrational amplitudes of the structural response will be
small. Conversely, if the energy dissipation is small, the vibrational amplitudes will I
I
increase as the absorbed energy is converted into kinetic or velocity energy (2.19) .
The load-deflection curve shown in Fig. 2-9b is called an elastoplastic re-

I
sponse, and is characterized by the flat plateau at the yield point (d-e). Generally,
structures wiIl not behave in a perfect elastoplastic manner. The exact behavior is
difficult to predict for structures, but it can be determined by tests for individual mem-
bers. Fig. 2-10 shows other common inelastic behavior patterns.
2.4.2 Displacement Ductility and Load Reduction Factors
I
Two methods are commonly used to equate the responses of elastic and I
I
elastoplastic systems. One method assumes equal maximum deformation, while the other

I
2.20
I
I
I
I
I
I
I
I
I (b)

I
Fig. 2-9 - I&spans of oecillntiorc3 to earthqueke motion
(a) Elastic responw. (b) Elestoplastic response. (Ref. 2.18)

I
I
1
I
I Py -- ------- 4’
1 Lc i
I i
I
4 Am.%

Fig. 2-10 - IdeaBed load deflection curves for common types of inelesticity

I
1
2.21
I
I
I
I
I
I
(a) (b)
Fig. 2-ll- Assumed responses of elastic and elastoplastic structures
(a) Equal maximum deflection response. (b) Equal maximum potential energy response.

method assumes equal maximum potential energy. Fig. 2-11 shows the responses for both
I
methods.
For both methods the ductility of the structure is measured by the displace-
I
I
ment ductility factor, u , defined as:

AU

I
v =*- (2-18)
Y

I
A, is the deformation at the end of the post elastic range, while A is the lateral
Y
deformation when yield is first reached.
The ratio of maximum inertia loads for the elastic to elastoplastic structures
is the load reduction factor R. For the assumption of equal maximum deformations, R is
defined as:
I
R=L
!J
(2-19) I
The SEAOC committee used this load reduction factor in determining appropriate code I
I
forces. They also indicated that typical values of u range from 3 to 5. As a result the
forces predicted by an elastic dynamic analysis will be 3 to 5 times larger than those
given by the code formula. Figure 2-12, shows this in graphical form (2.11) . ln theory, to
justify the lower design forces the structure must be capable of deflecting 3 to 5 times
the elastic deflection obtained with the code forces. I
I
Dynamic analyses conducted by Clough(2*20) have indicated the equal max-
imum deflection assumption is unconservative for degrading stiffness systems, such as

I
reinforced concrete. Fig. 2-13 developed by Clough suggests that for degrading systems

2.22
I
I
I
I
I
I
I
I
I
I I I / I
0 1.0 2.0 3.0 1.0 6.0

Pl)rl0d.T,**5

I
Fig. 2-12 - Comparison of equivalent static force method
to dynamic analyses procedures
the load factor defined by the equal potential energy assumption gives an upper bound to

I the system response. The load factor is obtained by making the area OCD, in Fig. Z-llb,
equal to the area OEFG. When these areas are equal the potential energy of each system

I
is equal and the load reduction factor becomes.

I
I
I
I
I
I Fig. 2-13 - Displacement ductility versus ratio of strength to elastic
demand for single degree of freedom oscillators responding to the 1940

I
El Centro N-S earthquake. (Ref. 2.18)

I
2.23
I
A comparison of lJ values obtained from the two methods is shown below for I
a range of R values.
I
lbble 2-3 Comparison of ductility factor, p

Design Load 0.2 0.4 0.6 0.8 1.0


I
I
R = Elastic response load

Equal maximum deflection u = l/R 5.0 2.5 1.67 1.25 1.0

Equal maximum potential energy, I


I
u = 1/2a2 + l/2 13.0 3.63 1.89 1.28 1.0

I
The comparison shows that for low strength levels the e&to-plastic system
will suffer larger displacements if it is to absorb the same energy as the elastic system.
It also shows that the difference between P values increases as the value of R
decreases. Thus for low values of R an exceptionally large amount of ductility may be
required. Although monolithically cast structures may be able to provide such ductility,
I
the capability of precast systems has yet to be established. Therefore, it may be appro-
priate to use conservative estimates of R and P in the design of precast structures. I
I
The ductility factors given in Table 2-3 illustrate how ductility requirements
vary with stiffness characteristics and the assumed elastic design loads. Generally in the

I
design of a structure, ductility requirements are implicitly satisfied by other provisions
such as reinforcement ratios and defIection limitations. Whether theses provisions are
adequate to satisfy the ductility demands of a given system is beyond the scope of this
report.
Until more specific provisions are developed, precast design should strive to
I
meet the ductility provisions of reinforced concrete. If the intent of these provisions
cannot be met, or if in the judgment of the designer a precast system has less ductility I
I
than a similar monolithic structure, a larger elastic design force may be appropriate for
the precast system.

I
2.4.3 Curvature Ductility
In obtaining structural ductility, separate ductility requirements are imposed

I
on the individual structural members. Since flexural deformations are dominant in most
members, member ductility is most commonly represented by a curvature ductility

I
factor, +u/$y , where @,, is the member curvature at the end of the post-elastic range

2.24
I
I
I and @y is the member curvature at first yield. ‘lhe magnitude of curvature ductility
required for a given member will vary with the structural shape, the location of the
I member, the plastic hinge length (i.e. the length over which inelastic behavior occurs)
and the yield mechanism of the structure.
I Shear and axial ductility factors can also be established, but inelastic behav-
ior is seldom justified for these deformational modes and elastic behavior is usually
I required throughout the total structural response.
For a frame structure, ductility can be obtained through inelastic column

I behavior, Fig. 2-14(a) or inelastic beam behavior, Fig. 2-14(b). Park and Paulay(2’18)
analyzed the frames of Fig. 2-14 and determined curvature ductility requirements for the
two mechanisms shown. ‘Iheir results indicated that for a ten story building with a
I displacement ductility factor of 4, the required curvature ductility factor with the
column sidesway mechanism was 122. For the beam sidesway mechanism the required
I curvature ductility was only 18.
This drastic difference, along with the impossibility of obtaining a curvature
I ductility of 122, illustrates why it is more desirable to form plastic hinges in the beams
and allow the columns to behave elastically. The beam-sidesway mechanism also results

I in a more stable structure with lower P-A moments and less permanent lateral deforma-
tion.
Fig. 2-15, developed by Parme(2.10) , also illustrates the effect of curvature
I ductility on inelastic structural response. Fig. 2-15(a) shows that formation of a hinge at
the first floor girder reduces the induced shear into the structure. Further reduction of
I the shear force requires formation of additional hinges and additional curvature ductility
from the first floor girder.

I
I ’/ Fi

I 0pc
w QPC

I
I E!a I

(a) Cdmn sidesway (b) Beam sidesway


Fig. 2-14 - Plastic hinge mechanism for a Qpicnl multi-story frame
I
I 2.25
0 I 1
01 2 4 6 a 10 12 14

b/by of first floor girder d/+$ of first floor girder

(a) (b)
Fig. 2-15 - helastic response of U-story building I

Fig. 2-15(b) shows how the ductility of the first floor girder wiI1 effect
required moment capacity of the base column. The figure shows if the beam behaves
the
I
elastically, +tI,/$ = 1 , the required moment capacity at the base would be four times
the elastic code value. However, if a curvature ductility of 14 is obtained, the capacity I
of the column need only be slightly greater than the code value.
These two examples of frame behavior illustrate the fact that there is a I
difference between structural displacement ductility and member curvature ductility. In
general, the curvature
resisting frames will
ductility
be from
demand for girders and connections of ductile
three to five times the required
moment-
displacement
I
ductililty(2’10).
For shear wall structures, ductility occurs with the formation of a plastic
I
hinge at the base of the shearwall,
curvature ductility ou/oy
as shown in Fig. 2-16.
, and displacement ductility
The relationship
Au/Ay , can theoretically
between
be I
calculated from the ratiofp/% (2.18). For example, for the structure in Fig. 2-16 the
following curvature ductility factors are required for a displacement ductility factor of I
four.
L,la
a$+,
0.05
20.5
0.1
10.5
0.15
7.2
0.20
5.6
0.25
4.6
0.30
3.9
0.35
3.5
I
For monolithic shearwalls I,, Is typically in the range of 0.5 to 1.0 times the
I
r

member depth.
connection
However, for low rise precast shear walls it ls not uncommon to have a
which does not develop the full flexural strength of the shear wall. If inelas- I
tic behavior is limited to the connection, the plastic hinge length is limited to the con-
I
2.26
I
I
I
I
I
I
Fig. 2-16 - Cantilever shear wall with lateral loading at ultimate moment
I nection length and considerably less than 0.5 times the wall depth.

I 2.5 Lateral Load Resisting Systems

The random motions of a building foundation during an earthquake causes


I forces to be generated within the structure. The lateral load resisting system distributes
these forces and transmits them to the ground. ‘lhe system must be capable of trans-

I mitting these loads within the limits of acceptable


tion being the safety of the building occupants.
damage and with primary considera-
The elements comprising these sytems

I usually function as vertical load carrying


being a part of the lateral load resisting system.
members or as space enclosures in addition to

Most precast concrete members, especially when they are prestressed, are
I factory-made products. Their economy stems from repetitive use of similar shapes and
construction details. A large project which can achieve repetition within itself may
I economically employ special products and details, but the vast majority
not fit this category. Thus, the designer of precast, prestressed concrete structures must
of buildings do

I learn to live within a discipline established by the manufacturing process. This concept,
as related to connections, is discussed in Sections 3.1.6 and 3.1.7.

I This section will concentrate on systems that can perform


resisting functions
the lateral
required of it and also use the precast concrete members and details
load

that are standard within, or easily adaptable to the manufacturing process.


I The basic lateral load resisting systems and components discussed in this
section are:
I 1.
2.
Precast shear walls
Precast ductile moment resisting frames

I
I 2.27
I
3. Braced frames
I
4. Floor and roof diaphragms
The function and performance of precast concrete lateral load resisting I
systems are largely dependent on the type of connection used. The illustrations in this
section show different types of connections which are designated schematically as shown
in Table 2-4.
I
Table 2-4 Schematic Connection Symbols
I
iymbol
A
Connection Type
Fixed connection. Stresses in the components remain in the elastic I
range throughout the earthquake. No significant movement occurs.
0 Ductile connection. Will yield or otherwise behave inelastically at high I
force levels, allowing energy dissipation.
0 Pinned connection. Allows relative movement when subjected to latera: I
loads.

2.5.1 Shear Wall Systems


I
The strength and stiffness associated with concrete shear walls has long been
recognized as a sound structural solution for resisting wind forces and deformations.
I
However for earthquake loading, the inherent stiffness at the shear wall compared to a
moment resisting frame has led building codes, such as the Uniform Building Code(2.15),
I
to impose higher design forces on shear walls than on ductile moment resisting frames.
This should not imply that a ductile moment resisting frame is a better solution for I
earthquake resistance than a shear wall system. Both systems, when properly designed,
can provide adequate resistance to strong earthquakes, just as both can provide adequate I
resistance to wind loading. In fact, there is much evidence that buildings having lateral
load resistance from shear walls exhibit much less structural and non-structural damage
than ductile moment frames during an earthquake (2.21).
I
Under the Uniform Building Code(2’15), precast shear walls may be designed
using one of three basic concepts:
I
1. Searing wall shear panels (K = 1.33)
2. Non-bearing shear panels with vertical loads supported directly by cast- I
in-place concrete sections or separate columns (K = 1.00).
3. Combined non-bearing shear panels with ductile moment frames capable I
of providing a minimum of 25% of the required seismic resisting capa-
I
2.28
I
I
I city. The shear wall panels must be designed for 100% of the required
seismic capacity, but the seismic design forces are reduced to allow for
I the resulting system redundancy by a reduced K factor (K = 0.80).
Boundary elements, as specified by Section 2627(c) of the Uniform
I Building Code, are required in the vertical edges of the wall panels in
Seismic Zones 3 and 4. Because of the difficulty in meeting these

I boundary conditions, it is usually not practical to use precast concrete


shear walls for this building category.
The higher K factor assigned to bearing wall shear panels increases the re-
I quired elastic behavior of the wall, but the 1.33 value is insufficient to eliminate over-
stress in the system during a major earthquake. Ductility and energy dissipation is still
I required with the forces obtained through design codes.
The inelastic behavior and ductility of a shear wall will vary with the size,
I shape and construction of the wall. Short shear walls with a height to length ratio less
than one will behave like a deep beam with shear deformation contributing a significant

I portion of the total structural response.


Energy dissipation and inelastic deformation in short shear walls is very
difficult to obtain. This is due to the presence of low flexural stresses and high shear
I stresses. If inelastic deformation is to occur it would most likely be from shear sliding of
the base(2’22). Later sections will show, however, that for most connections inelastic
I displacements associated with sliding shear can significantly reduce the energy dissipa-
tion of a connection with progressive cyclic loading. Hence, for short shear walls it is
I necessary to increase the design loads and allow for more elastic behavior instead of
relying on inelastic shear sliding. However, strength degradation and ductility must still

I be considered since a large number of load reversals will occur with the low periods
associated with short shear walls.
As the ratio of height to length increases, the deformation characteristics of
I the shear wall change from a deep beam to a cantilever beam. This reduces the influ-
ence of shear deformation and increases the contribution of flexural deformation. ln-
I elastic deformation is then obtainable through yielding of the flexural reinforcement and
plastic hinge formation as described in Sec. 2.4.3.
I gories:
Precast concrete shear wall systems can be divided into the following cate-

I 1.
2.
Single panel - isolated
Single panel - coupled
3. Stacked panel - isolated
I
I 2.29
I
4. Stacked panel - coupled
I
The type of joinery between the vertical and horizontal boundaries of the
individual wall elements will determine the type of system, and whether the elements I
should be designed as individual units standing; alone (isolated), or as part of a larger
shear element (coupled). These concepts are illustrated schematically in Figs. 2-17 I
through 2-22.
Connections are required to:
1. Transfer shear from the floor and roof diaphragms to the walls.
I
2.
3.
Transfer shear from the walls to the foundations.
In stacked panel systems, transfer horizontal shear between panels.
I
4. In coupled systems, transfer vertical shear between panels.
5. Transfer tension caused by overturning forces. I
2.5.1.1 Single panel systems (Figs. 2-17 and 2-18)
I
Here, there are no horizontal joints between precast panels. Lateral shears
are carried directly from the floor and roof diaphragms through the wall to the support-
ing foundation.
I
This type of wall is frequently used in single-story industrial buildings and low
to medium rise commercial and residential structures, most often as exterior walls. I
Panels are cast in long-line forms and are usually prestressed. They may be either
stemmed units, such as double tees, or flat panels and are often insulated as sandwich I
panels. The height of the panel is only limited by the transporting and handling consider-
ations. I
2.5.1.1.1 Single Panel-Isolated
When single panels are isolated, as shown in Fig. 2-17, there is no connection I
between units. The stiffness of the wall system is equal to the sum of the stiffnesses of
the individual panels and the lateral load distributed to each individual wall is propor- I
tional to its shear and flexural stiffness.
The base connection of an isolated shear wall panel can be designed for
elastic or inelastic behavior. If the connection is to behave elastically one of two condi-
I
tions must be satisfied. Either the connection must be designed for the maximum earth-
quake loads, which may be four or five times the equivalent static load, or the capacity
I
of the connection must be greater than the capacity of the wall. The latter case assumes
formation of a plastic hinge in the wall thereby limiting the loads induced into the con- I
nection. The performance of the system is then dependent on the ductility and energy
I
2.30
I
I
I
I Root --- I
I Floor
El
---
I n
I Floor ---
I Floor
1
--- ---

I
I Fig. 2-17 - Jsolated single panel shear wall system

I
I
I
I
I
I
---- --- ---
<~t,~L/L~=Az-~~ -- I

I
Fixed Coupling Ductile Coupling

I
Fig. 2-18 - Coupled shgle panel shear wall system
I 2.31

I
I
Illfill Ppr:3!,_lnfill
l--twall width wall
1 I
I
I
I
I
I
I
Fig. 2-19 - Jsolated stacked panel shear wall system I
I
I
I
I
I
I
I
W- 2-20 - Coupled stacked panel shear wall system 1. Fixed couplings.
I
I
2.32
I
I
I
I
I
I
I
I Fixed coupling

I pig. a-21- Coupled stacked panel shear wall system. Jntermittent ductile couplings.

I Panel length

I -

I Z=
rii
0”s
I
I
I
I
I.
. . * 4.
I rnt
II -

I
Fig. 2-22 - Coupled stacked panel shear wall system. Duetile coupli~.
I
I 2.33
I
absorption of the wall instead of the connection. I
If the connection can develop sufficient inelastic behavior, lower elastic
deskn forces, such as those predicted by an equivalent static method, may be used for I
Proportioning the connection components. The inelastic behavior of the connection
should ahow rotational ductility to develop at the base of the wall before a brittle shear
or concrete failure occurs in the wall or connection.
I
For example, consider the single story shear wall and base connection shown
in Fig. Z-23. The connection forces, C and T, are determined by the wall dimensions, the
I
lateral load, V, and the axial load, P. During an earthquake the magnitude of V will
increase until the maximum acceleration is reached, or until yielding occurs in the wall I
or connection.
For the connection shown in Fig. 2-23(a), the preferable inelastic behavior I
will occur with the yielding of the reinforcing bar. This will produce the longest plastic
hinge length and the greatest rotational ductility at the base of the wall. Yielding of the
connecting plate can also produce rotational ductility, but the short length of the plate
I
reduces the plastic hinge length and increases the ductility demand as described in
Sec. 2.4.3. Yielding of the embedded studs (Fig. 2-23(b)) is dependent on the brittle
I
fusion weld attaching the stud to the plate and therefore is not recommended.
If the reinforcing bar is to provide the ductility of the connection it should be I
the first component to yield. At yield the magnitude of the lateral force, V, will remain
constant until strain hardening initiates in the rebar as shown in Fig. Z-24. For grade 60
steel, strain hardening will allow the force in the bar to double the yield force. Conse-
I
quently, the maximum value of V may be twice the value which initiates yield in the
bar. If all the components of the connection, except the reinforcing bar, are designed
I
I
I
I
I
t i Ial lb) I
Fig. 2-23 - kolated shear wall and base connection
I
2.34
I
I
I
I
I
I
I 0
0 0.04 0.08 0.12 0.10
0.200
J

S,,d"

I Fig. 2-24 - ‘Apical stress-strain diagram for reinforcing bars

I elastically for two times the lateral force which initiates yield, the bar and connection
will obtain maximum ductility.

I Connections utilizing headed studs in both connecting parts, as shown in


Fig. 2-23(b), are very brittle. The tension force, T, for such a connection is developed by
shear through the studs in the panel. This mechanism possesseslittle ductility and con-
I nections of this type should be designed elastically for the maximum earthquake force.
If the axial load acting on the panel is sufficiently large, or if the height to
I width ratio is small, tension may never develop in either connection. When this occurs
the wall and its connections will behave elastically with the maximum shear induced into
I the wall occurring with the maximum amplified acceleration.

2.5.1.1.2 Single Panel-Coupled


I Adjacent panels of coupled systems, Fig. 2-18, may be coupled in one of two
ways:
I 1. Rigid connections such that the composite wall possesses relatively
little capability to deform inelastically in a maximum credible earth-
I 2.
quake.
“Flexible” links between vertical elements, such that the interconnected

I members behave compositely during minor seismic disturbances, but the


connections deform, or slip during a major earthquake and thereby
dissipate energy.
I The intent of coupling precast panels together is to produce a stiffer shear
wall system, and to reduce the overturning stresses in each individual panel. For in-
I
2.35
I
I
I
I
I
I
Fig. 2-25 - Behavior comparison of coupled and uncoupled shear wall systems I
stance, the three coupled panels shown in Fig. 2-25 will have greater overturning resis-
tance and leas lateral deformation than the three uncoupled panels subjected to the same
I
force.
When the coupling connections are flexible or degrading, the coupled system I
obtains the characteristics of an uncoupled system after the connections yield or de-
grade. ‘Ibis increases the flexural stresses at the base of the wall which induces larger I
forces into the base connection. These larger forces must be accounted for in the design
of the base connection by either increasing the elastic design forces, or by providing
sufficient ductility in the connection.
I
2.5.1.2 Stacked Panel Systems (Figs. 2-19 through 2-22) I
(Note: Sometimes called Urge panel” systems)
Precast concrete wall elements are stacked vertically and interconnected in I
the field. Although studies have shown that flexible links can be provided through the
horizontal joints, it is seldom practical in taller structures. The horizontal joint
connection may be directly from panel to panel, or it may be through the floor system.
I
This type of wall is most frequently used in mid-to-high rise commercial and
residential structures. They are used for interior and exterior, load-bearing and non-load
I
bearing walls. The panels are usually cast in individual forms and not prestressed in the
plant, although they are sometimes post-tensioned together vertically as a connection I
method.
Coupled walls may use either rigid or flexible links. Research is currently
being conducted on energy absorbing connections for the vertical joints between panels.
I
III addition to walls coupled in the same plane, as shown in Fig. 2-19 through
2-22, they may be coupled at right angles to form L-, T-, or box shaped sections, adding
I
rigidity.
I
2.36
I
2.5.2 Precast Frame Systems
The Uniform Building Code requires that all buildings utilizing a “K” factor of
0.67 or 0.80 have a ductile moment resisting frame. All buildings over 160 feet in height,
except in Zones 1 and 2 with a design K = 1.00 or 1.33, must have a ductile moment
resisting frame, either to take the total seismic loading, or at least 25% of the seismic
loading when a shear wall system is used for 100% of the seismic load. Seismic frames
for seismic Zone 1 locations are exempted from the ductility requirements and are not
required in buildings exceeding 160 feet in height.
Ductile moment resisting frames are difficult to achieve with totally precast
concrete structural systems. More frequently, composite systems, with cast-in-place
reinforced concrete handling the burden of the moment connections offer superior solu-
tions. In many cases, when conditions require the use of ductile frames, it is more eco-
nomical to consider alternate structural systems, such as reinforced cast-in-place con-
crete or steel, for the frame and using precast for just the deck members.
When composite frames are used, ductility is usually achieved by lapping or
splicing reinforcing bars within the connection. Post-tensioning may also be used to
connect beams to columns in a ductile moment-resisting frame. While this is not pre-
sently covered by the Uniform Building Code, extensive research in New Zealand indi-
cates there is excellent energy dissipation in precast framing members connected by
post-tensioning. Examples of both reinforced and post-tensioned connections are shown
in Part 4.
Several successful projects have used special sections such as “cruciform” or
“tree” precast combined beam/column sections, jointed at points of low flexural stress
(See Fig. 2-26). While these elements meet the requirements for ductile frames, and
often provide interesting architectural features, their fabrication does not adapt well to
the manufacturing process. As such, they are only economical on large structures where
repetition can be achieved within the project-much the same as architectural precast
wall panels.
When Precast, prestressed members are connected with moment-resisting
connections, the effect of restraint of shortening from creep, shrinkage and temperature
change must be a primary consideration in design. The build-up of these forces can be
minimized by strategic location of connections which will accommodate movement. This
is discussed more completely in Part 3.
2.5.3 Braced Frames

Diagonal steel bracing is often employed in structural steel frames for reduc-

2.37
I
I
I
I
I
Fig. 2-26 - Examples of combined precast beam/column mtion.q
I
Used in ductile moment+esktii frames.
ing lateral forces caused by wind or earthquake. Three types of framing bracing are I
shown in Fig. 2-27.
The K-brace and the X-brace systems are the two most familiar systems
which have been investigated and tested for earthquake loading. Both systems initially
I
offer considerable strength and stiffness, but inelastic lengthening and cyclic bucking
usually reduces the strength and stiffness of the bracing members. Thus for proper
I
inelastic cyclic analysis of braced frames, an accurate hysteretic model for individual
braces must be available(2’68). I
The eccentric brace system, shown in Fig. 2-27(c), is a relatively new system
which has shown promise in steel construction. The system is unique in that it has the I
strength and stiffness of a braced frame with the inelastic behavior and energy dissipa-
tion of a moment-resisting frame (2.30).
The only research available on eccentric bracing involves the behavior of a
I
steel system. The inelastic action observed in the frame was obtained through shear
I
I
I
I
a) x-erase b) K-Brace c) Eccentric
smc* I
Fig. 2-27 - Cross bracing used with steel frames
I
2.38
I
I
I yielding in the web of the “link beam” between the column and the brace. This type of
yielding cannot happen in precast concrete members, so additional analytical and labora-
I tory research would be required if such a system is to be used with precast concrete
frames.
I Little if any construction has been done using braced precast concrete frames
for earthquake-resistant structures. However, braced systems have been utilized to

I resist wind loads(2’31), usually with steel sections used for the bracing members.
The advantages of braced frames include:

I 1. Standard rectilinear sections can be used for the precast beams and
columns.
2. Economical “pinned” connections, which erect rapidly can be used at the
I beam column interface.
3. Volume change restraint forces are reduced.
I 4. Erection is not slowed by time-consuming moment connections or cast-
in-place infill shear walls.
I The UBC recognizes braced concrete frames and requires the connections of
the system to be designed for the capacity of the member or 1.25 times the equivalent

I static force.
For braced steel frames over two stories in height ATC-3 requires the mem-
bers to have a compressive strength equal to at least 50 percent of the required tensile
I strength. ‘This is due to the poor earthquake performance of braced frames which con-
tained only tension members. Compliance with this provision would be proper for con-
I crete frames utilizing steel bracing members.
2.5.4 Floor and Roof Diaphragms
I The diaphragm floor and roof systems of precast structures are composed of
carrying girders, beams, deck elements and often a cast-in-place concrete topping slab.
I These are designed initially to accommodate the vertical load requirements. ‘lhe con-
nection details are designed to transfer the vertical loads between the various elements
I and to account for the member induced effects of shrinkage, creep and temperature
variation.

I As part of the lateral load resisting system, the diaphragm is designed as a


relatively rigid horizontal beam to distribute the lateral loads to the various shear wall
or frame elements.
I The lateral loads to be used for the design of a given diaphragm will vary
from level to level depending on its location vertically in the building. The design load is
I
2.39
I
I
based primarily on the maximum horizontal acceleration that is calculated for that
I
particular level in the building, with minimum values prescribed by Code. Changes in
stiffness of shear elements directly above and below the diaphragm must be considered in I
the design.
The loads assumed by the Codes for diaphragm design are rather I
approximate. Such things as openings and frame or shear wall stiffnesses have an effect
on the distribution of loads. ‘lhus, design refinements are not warranted, and simpJ,ifying
design assumptions are normally used.
I
Fig. 2-28 shows three general design cases for diaphragms.
The following elements should be considered in the analysis and design of
I
horizontal diaphragms:
1. There must be sufficient shear strength between members to act as an
I
integral structural beam web; enough perimeter (flange) reinforcement
to develop flexural tension; and adequate connections to transfer shear I
to the shear walls or frames. Perimeter bars in the frame or shear wall
may be used as flange reinforcement.
Struts may be required between shear walls or frames to distribute
I
2.
forces to the diaphragm, to transfer forces between vertical shear
elements, or to redistribute forces around openings in the floor dia-
I
phragm. Cast-in-place concrete pour strips may be used for this pur-
pose if mechanical fasteners cannot handle the forces.
I
3. The diaphragm is usually assumed to be an infinitely stiff element for
analysis of the load distribution between the shear elements. If this is I
not the case the relative stiffness of the diaphragm must be
considered. Diaphragm deflections are normally very small and most
are caused by shear &formation rather than flexural deformation.
I
However, if the diaphragm has substantial flexibility these deflections
should be added to the story drifts contributed by the shear frames or
I
4.
walls.
Horizontal torsion resulting from the center of mass being eccentric to
I
the center of rigidity must be included in the design. Most codes re-
quire that five percent of the longest plan dimension is the minimum I
eccentricity to be considered.
5. Shear transfer of the diaphragm with topping is dependent on the bond
of the topping to the precast element. Diaphragm systems without
I
I
2.40
I
I
I d
I
w’l 1
I
F,= Wd
I R,=R2=Fx/2
T=C=F.d
I4 xii

I
I
I (a) Simple beam action, lateral suppoh at ends

I
I
I Y = .F, = d/? tF,l
R3=R,=dF./2h

I T=C=(M-wd,2/2)
h

I
0 Cantllever beam action, lateral suppwts at thref2 sides
I
I
I
I
I
I
I
I
I
topping must transfer these generated forces through mechanical con-
I
nections or continuous pour strips. Mechanical connections cause stress
concentrations which may result in local failure under the possible peak I
load conditions. This may be avoided by reducing the allowable stresses
used for design of the connections and to thus increase the elastic
capacity. Continuous pour strip connections result in a more uniform
I
6.
stress distribution, but are much more costly.
Interior bearing walls and slab openings tend to disrupt diaphragm
I
7.
continuity. Additional reinforcement may be required in these areas.
Earthquake loads may be induced into a diaphragm from any direction.
I
The diaphragm and its connections should be designed for the most
critical loading direction. For convenience ATC-3 requires 100 percent I
of the forces for one principal axis plus 30 percent of the forces for the
perpendicular axis. This combination is considered for both principal I
directions and the combination requiring the maximum component
strength is used. I
2.6 Dynamic Characteristics of Precast Connections
Part 3 of this report discusses various design principles and behavior char- I
acteristics for precast connections subjected to monotonic loading. During an earth-
quake, connections are subjected to monotonic gravity loads plus time-varying cyclic
lateral loads. The behavior of a connection during cyclic loading can be different than
I
for static loading and as the previous sections have shown, the behavior of the connec-
tions will influence the structural reponse and the loads induced into the structure.
I
Properties such as initial rigidity, ductility, inelastic deformation, strength and stiffness
degradation as well as the loading influence connection behavior and structural response
I
to earthquake loads.
This section describes some of the characteristics associated with typical I
precast connections. The limited dynamic experimental data which is available is pre-
sented and areas in need of future research are noted. Although design of precast con- I
nections for seismic loading is based on monotonic elastic behavior, knowledge of the
cyclic behavior of a connection is essential for preventing unwanted brittle failures and
for accurately predicting the seismic response of the precast system.
I
2.6.1 -Shear Connections I
Shear connections are classified as either “wet” or “dry”. Wet connections
I
2.42
I
I
I
I a) my Conn*cthm b) Wet Connection

Fig. a-29 - sear conne.ctions


I use reinforced or unreinforced cast-in-place concrete to form the junction between
members. Dry connections utilize a mechanical anchor, such as bolts or welded metal, to
I transfer the load. An example of each connection is shown in Fig. 2-29.

I 2.6.1.1 Wet Connections


The significant properties which influence the strength and performance of

I wet connections include the precast and cast-in-place concrete,the amount of transverse
and longitudinal steel, the tensile or compressive forces acting on the connection and the
surface preparation of the precast panels. Fig. 2-30 shows some of the typical edges
I presently used in precast construction.
Shear transfer in wet connections can occur through cohesion or bond, fric-
I tion, dowel action and direct bearing. However, the construction process, along with
creep and shrinkage effects, will often destroy the bond between precast and cast-in-
I place concrete thereby eliminating that transfer mechanism(2.32) .
Shear transfer through friction can only occur with a clamping force acting

I normal to the connection face. The clamping force can be supplied through external
compressive forces, post-tensioning, or transverse mild steeL When transverse mild steel
is used the shear friction principles discussed in Sec. 3.3.2.3 are used to determine the
I magnitude of shear transfer.
“Effective” shear-friction, as discussed in Sec. 3.3.2.3 was developed from the
I results of monotonic tests. During cyclic loading, tests have indicated that these princi-
ples may be unconservative and that a reduction in shear transfer will occur during
I successive load cycles. Mattock obtained this result in his cyclic shear transfer testing

I
I
I
Fig. 2-30 - Typii panel Bage prepsration
I
I 2.43
I
I
I
I
I
I
I
(a) cycles l-5 b) cycks 25-30
Fig. 2-31 - sheer-slip from Univ. of Washington teats (Ref. 2.33) I
at the University of Washington. Hawkins in his state-of-the-art report (2.33), provides a I
summary of Mattock’s work.
A typical load-deformation curve for one of Mattock’s test specimens is
shown in Fig. 2-31. Fig. 2-31(a) shows the reponse for the first five cycles at a load of
I
0.5 V”. During this loading the specimen was stable with little loss of stiffness or
strength occurring. However, as the load and number of cycles increased, the stiffness
I
and strength of the connection degraded. This is evident by the increase in slip between
cycles 25 and 30, Fig. 2-31(b). I
In comparing the cyclic results against monotonic load results, Mattock
observed a reduction in strength of about 20% in the cyclic tests for initial crack widths I
of 0.015 in. or less. The reduction increased with increasing initial crack width and was
more for lightweight concrete than for sand and gravel concrete (2.33).
The energy dissipation capacity of the connection, represented by the area
I
under the shear-slip curve, also degraded with cycling and increasing load. Tests on
shear wall structures(2*34) have indicated that under axial loads the hysteresis loops
I
become fuller, but in general shear transfer should not be relied on to dissipate much
energy. I
Development of transverse steel beyond both sides of the failure plane is
essential for wet connections. For some precast connections, such as the hollow-core I
connection shown in Fig. 2-32(a), this development is easily obtained.
I
2.44
I
I
I
I
I (b) (C)

I Fig.2-3%-slabtoslabcoMections

A straight lap connection, as shown in Fig. 2-32(b), is not recommended for

I seismic regions since the lap length is insufficient for most joint widths. Welding, me-
chanical connectors, or longitudinal steel through hooped bars, Fig. 2-32(c) are possible

I solutions.
2.6.1.2 Dry Connections

I The most common type of dry shear connectors are composed of embedded
steel shapes anchored into the precast members by studs or reinforcing bars. The con-

I nection is completed by bolting or welding a third element to the embedded steel


shapes. Fig. 2-33 details some of the most common dry connections.
Unlike wet connections, where shear friction is the main mechanism of shear
I transfer, dry connections involve a variety of mechanisms in developing shear resis-
tance. Depending on the detail, shear can be transferred through bearing of the steel
I shapes, shear of the connecting elements, shear of the welds or bolts, or through friction
between bolted plates. If the space between the panels is grouted, shear transfer through
I shear friction can also be developed, but then the dry connection details are subjected to
forces associated with clamping rather than shearing.

I Many tests have been conducted on the ultimate strength Of dry COMedOnS
under static or monotonic loading, but few tests have been conducted under cyclic or
alternating loads. Spencer and Neille(2’35) are two of the few who have conducted and
I reported on cyclic tests. Their work at the University of British Columbia involved the
cyclic testing of the six specimens shown in Fig. 2-34.
I Five of the test specimens were subjected to cyclic loading at frequencies in

I
I (b) (Cl
Fig.2-33-Dtysbefueonnecto~
I
I 2.45
I
CONNECTIOh I
DETAILS
PANEL
OF STUDS 8
REINFORCEMENT I
I
Al
I
A2.A3
I
I
81
I
B2
I
83 I
I
pig. 2-34 - Details of test connectiom

the range of 0.01 to 0.02 cycles per second. Connection Al was loaded monotonically to I
failure. The load was applied ‘l/8 inch from the face of the angle. This produced tensile
and compressive forces in the studs equal to 10 and 15 percent of the shear force for 12 I
and 10 inch long connection angles, respectively.
Test results for connections Al and A3 are shown in Fig. 2-35. Fig. 2-35(a)
shows that connection Al exhibited considerable ductility under monotonic load. The
I
yield strength of the connection was greater than the design strength of 27.2 kips, pre-
dicted by the PC1 Design Handbook(2.36) .
I
The cyclic test results for connection A3, depicted in Fig. 2-35(b), illustrates
the change in behavior between monotonic and cyclic loading of the same connection. I
The connection was loaded to yield after several initial cycles in the elastic range, and
then subjected to steadily increasing deflections. The results again show a decrease in I
stiffness and strength with cycling.
8pencer and Neille reported that failure of the connections were preceded by
progressive crushing and spalling of the concrete above and below the angle. They also
I
observed progressive tension cracking parallel to, but several inches away from the
angle. Connections Al and Bl failed when a large block of concrete surrounding the
I
I
2.46
I
I
I
I
I
I
I +I / “El0ENYELOFcy
1L30
Load-deflection curve bnnection Al) for
I monotonic load* to failure (b) Load-deflection loops (connection A3)

Fig. 2-35 - Load deflection curves for connections shown in Pi. 2-34 (Ref. 2.35)
I studs broke away. The other connections failed when one of the studs broke close to the
I fusion weld.
The following conclusions were reached by Spencer and Neille after the cyclic

I testing(2’35).
1. The PC1 procedures for calculating the ultimate design strength of
these connections under static loading give conservative results.
I 2. ‘lhe strength of the connections in the first cycles of loading up to yield
will be approximately the same as the strength in mon~otonicloading.
I 3. If cyclic loading is continued above the stability limit, the strength of
the connections will fall with an increasing number of cycles, and the

I 4.
yield strength envelope will tend to approach the stability limit.
The deflections reached before failure were from seven to twenty-four

I times the theoretical elastic deflection corresponding to the design


ultimate strength.
5. These connections, if properly designed and detailed, appear to be
I suitable for use in earthquake-resistant buildings designed as box-type
systems.
I The fact that the weld between the stud and the angle failed in four of the
connectors suggests that the fusion weld is a weak point in the connection. Hence, if

I studs are to be used in a connection adequate quality control should be employed to


ensure a quality weld is produced and additional studs should be added to compensate for
imperfections. Hawkins suggests fifty percent of the studs should be capable of carrying
I
I 2.47
I
I
I
I
I
I
pip. !Z-36- Anem& embedment
I
the maximum load on the unit(2*33).
An alternate method for embedding steel sections in precast members is to
I
replace the headed studs with welded reinforcement as shown in Fig. 2-36.
Although sufficient seismic test data is not available, there are several
I
advantages to this type of connection. The first advantage is the larger surface area
available for welding the reinforcing bars to the angle. This allows for lower weld I
stresses and reduces the chance for a brittle weld failure.
The second advantage of the connection is the longer development of the bars I
into the precast member, which reduces the chance for pull out of the connection.
In general, dry connections should be designed conservatively for earthquake-
resistant construction. Longer weld lengths, additional anchorage of reinforcing bars and
I
lower allowable stresses are justified for cyclic loading especially with the presence of
forces normal to the shear connection. Reliminary testing(267) has shown that the
I
application of a one kip pull out force normal to a connection, similar to the one shown in
Fig. 2-33(b), reduced the ultimate shear capacity by l/3. Normal forces of this magni- I
tude can easily be produced at erection time, or from the effects of shrinkage and tem-
perature. Thus a lower rated capacity is warranted, especially in high seismic regions. I
2.6.2 Beam-Column Connections
Beam-column connections can be classified as rigid, semi-rigid, or simply I
supported. Rigid connections are incorporated into frame structures to provide resis-
tance against vertical and lateral loads. They are capable of transmitting shear, mo- I
ment, and axial loads. Semi-rigid connections are similar to rigid connections except the
I
2.48
I
I
I
I
I
I
I (a) W
Fig. 2-37 - Response of frame with pinned connections

I magnitude of moment transfer is reduced by a lower yield moment. Simply supported


connections allow unrestrained rotation of the beam member and transfer only shear and
I axial forces.
Semi-rigid and simply supported connections are usually designed to resist

I only gravity
to accommodate
loads. Their performance
the rotational
during an earthquake is dependent on their ability
deformations imposed by the structure. For simply

I supported connections, rotational deformations


for semi-rigid connections excessive rotation
produce little if any concern. However,
may induce stress concentrations into
portions of the connection resulting in unexpected connection or member failure.
I For example, consider the pinned frame shown in Fig. 2-37. Theoretically the
frame cannot resist lateral loads. Its function is to resist vertical gravity loads and to
I transfer lateral loads to other suitable portions of the structure. During an earthquake
the frame would undergo lateral deformations as shown in Fig. 2-37(b). The magnitude of

I deformation would depend on the structural shape, the lateral load resisting system, the
diaphragm stiffness and the physical properties of the frame. As a result of the lateral

I frame deformation, moments and shears are induced into the columns by P- Aeffects and
large rotations occur at the beam-column connection. The detailing of the pinned con-
nection must be able to accommodate these deformations and forces.
I Two common types of simply supported precast connections are shown in
Fig. Z-38. The connection shown in Fig. 2-38(a) has a large rotational capacity which
I could accommodate the deformations of a major earthquake with little distress.
See Part 4 for other features of these connections.)
(Note:

I In Fig. 2-38(b) an angle is added to the top of the beam for lateral support.
The connection is not designed to transmit moment, but rotation of the beam will induce
stress into the angle and the adjoining parts. Proper design of the connection requires a
I
I 2.49
I
I
I
I
I
I
(b)
I
Fig. 2-38 - Typical pinned amnections I
flexible angle at the top of the beam so that the angle will deform before large stresses
are induced into the connection. If the angle is too stiff, or welded along both sides,
large rotations which force the bottom of the beam to bear against the column could
I
produce failure in the top connection.
The consequences of local failures in pinned connections may seem minor, but
I
redistribution of load or damage to the column or beam could effect the structural
performance significantly. I
Rigid beam-column connections for ductile earthquake resistant moment
frames should conform to the following provisions: I
1. The sum of the moment strengths of columns at factored axial loads
should be greater than the sum of the moment strengths of the flexural
members along each principal plane. This results in the preferred
I
beamsidesway failure mechanism instead of the column-sidesway
mechanism as shown in Fig. Z-14.
I
The curvature ductility capacity of the connection must be sufficient to
2.
accommodate the ductility demands imposed by the inelastic structural I
response. Methods for determining the curvature ductility required of
members designed by an equivalent static method have yet to be estab- I
lished. This is a result of the complex nature of frame structures and
the many variables which influence the formation of the structural yield
mechanism. Generally, for ductile reinforced concrete frames a curva-
I
ture ductility of three times the lateral ductility is acceptable.
I
2.50
I
I
I 3. The shear capacity of the connection must be greater than the flexural
overcapacity of the beam. This ensures the formation of a plastic hinge
I and ductility in the connection before the occurrence of a shear failure.
2.6.2.1 “Dry” Beam-Column Connections
I Moment resisting connections utilizing precast members and mechanical
anchors are commonly used to resist wind or gravity loads. However their use in resist-
I ing lateral seismic loads has been limited by the lack of test data concerning cyclic
strength, stiffness and ductility. Without further testing the acceptance of these con-
I nections will be limited to areas of low seismic activity.
Pillia and Kirkt2’3g) are currently conducting a series of tests on the cyclic
I behavior of the precast connection shown in Fig. 2-39. The load history for the testing is
shown in Fig. 2-40 and the moment rotation curve for one specimen is given in Fig. 2-41.

I Failure in six out of seven test specimens occurred by fracture of the weld
connecting the bottom beam reinforcement to the angle. Inspection of these welds after
fracture showed that they were somewhat incomplete. Tnis emphasizes the importance
I of quality workmanship in welded connections and suggests that the connection can be
improved if a better transfer mechanism can be obtained between the reinforcing and the
I angle.
Failure of the precast connections oecured between the seventh and fifteenth
I cycle. Loading of one specimen was terminated in the fourteenth cycle when the capa-
city of the loading jack was reached. A similar monolithic connection was tested and

I failed after the seventh cycle by buckling of the bottom bars.


Reference 2.39 is a preliminary report since only eight of a planned eleven

I
I
I
I
I
Fig. 2-39 - Test specimen we-d in FWia-Kirk tests (Ref. 2.39)
I
I 2.51
I
I
I
I
I
I
%- GO - Loade pattern used in Pillia-Kirk tests (Bf. 2.39) I
tests were conducted at the time the report was published. However, the initial test
results are encouraging and the performance of these specimens provides optimistic hope I
for the development of similar connections.

2.6.2.2 Cast-In-Place Joints


I
will utilize
Until further
cast-in-place
testing is conducted the majority
concrete to form the joint
of beam-column
between the precast
connections
members.
I
Currently this is the most reliable
resisting frames.
means of using precast members in ductile moment
I
I.2
I
I
I
I
I
I
Fig. 2-41- Typical momentmtfdion curve (Ref. 2.39)
I
2.52
I
I
I Obtaining continuity between the precast and cast-in-place reinforcement

I can occur with lap splices, mechanical connectors or welding.


The following ATC-3(2*16) provisions are required for lap splices and are
typical of most earthquake-resistant building code requirements.
I 1. Lap splicing of tensile reinforcement is permitted only if hoop or spiral
reinforcement is provided over the lap length. The maximum spacing of
I the transverse reinforcement over the length of the splice should not
exceed the smaller of 4 inches or d/4.

I 2. Lap splices should not be located (1) within beam column joints, (2)
within a distance of twice the member depth from the face of the joint,

I or where flexural yielding may occur with inelastic lateral displace-


ments of the frame.
3. Lap splices in columns are permitted only within the center half of the
I column span.
Welded and mechanical splices are permitted by building codes provided not
I more than alternate longitudinal bars in beams or columns are spliced at a section and
the distance between splices is 24 inches or more. However, the commentary to

I ~~&2.16) states that this requirement is related to construction convenience and


serviceability rather than to strength. If this is true, the code provision is defeating its
purpose with precast construction since the most convenient splicing of precast members
I is through one plane section.
The acceptance of specific mechanical couplers will vary with location and
I building authority. There are certain couplers which have been tested under cyclic loads
and accepted by building authorities throughout many parts of the world. Use of
I mechanical couplers in high seismic regions without appropriate cyclic performance data
is not recommended.

I 2.6.2.3 Post-Tensioned Joints


Beam-column connections can also be developed by post-tensioning the beam
I to the column. Although American building codes have yet to encourage the use of post-
tensioned connections, their acceptance is growing in other parts of the world. An

I interim report by the Commission on Seismic Structures of the FIP(2S47)encourages the


use of post-tensioned beam-column joints. Several such connections are shown in Part 4.

I The Standards Association of New Zealand has drafted seismic design recom-
mendations for prestressed concrete (2.45, 2.46) which includes provisions for post-ten-
sioned beam-column joints. Park, in Reference 2.41, describes some of the code re-
I
2.53
I
I
quirements and the research which led to their development. Several provisions listed by I
Park are given below.
“22.6 Joints in Prestressed Frames I
22.6.1 - Anchorages for post-tensioned tendons shall not be placed within
beam-column joint cores.
22.6.2 - Except as provided by Section 22.6.3, the beam prestressing
I
tendons which pass through joint cores shall be placed at the face of the
columns so that at least one tendon is located at not more than 6 in. (150 mm)
I
from the beam top, at least one within the middle third of the beam depth
and at least one at not more than 6 in. (150 mm) from the beam bottom. I
22.6.3 - When partially prestressed beams are designed with mild steel
reinforcement providing at least 80% of the seismic resistance, prestress may I
be provided by one or more tendons passing through the joint core and located
within the middle third of the beam depth, at the face of the column. In such
cases post-tensioned tendons may be ungrouted, provided anchorages are
I
detailed to ensure that anchorage failure, or tendon detensioning, cannot
occur under seismic loads.
I
22.6.4 - Ducts for post-tensioned grouted tendons through beam-column
joints shall be corrugated, or provide equivalent bond characteristics. I
22.6.5 - Connections between precast members at beam-column joints
shall be acceptable brovided that the jointing material has sufficient strength I
to withstand the compressive and transverse forces to which it may be sub-
jected. The interfaces shall be roughened or keyed to ensure good shear
transfer and the retention of the jointing material after cracking.”
I
2.7 Future Testing I
A great deal more seismic testing of connections and components is probably
essential if precast prestressed concrete is to develop into a generally accepted safe I
system for earthquake-resistant design. Without additional research and testing to
determine appropriate dynamic characteristics, precast and prestressed systems will be
subjected to design uncertainties and strict code requirements.
I
If rapid advancement is to occur in the acceptance of precast systems proper
organization and standardization of test procedures will be required. Presently there is
I
uncertainty as to what constitutes an appropriate test, what demands are imposed On
connections and how to determine if a connection is acceptable for seismic loading. I
Although there are no general answers to these questions, guidelines could be established
I
2.54
I
I
I to eliminate much of the confusion which now exists.
In establishing the demands imposed on a system the following parameters are
I important(2’4g).
1. Maximum amplitude of deformation
I 2.
3.
Number of large-amplitude cycles of deformation
Sequence~oflarge-amplitude cycles of deformation
I 4. Associated maximum forces
Developing representative values for these parameters will allow easier evaluation of

I performance data and reduce skepticism concerning acceptable performance behavior.


Research has been conducted to develop appropriate loading programs for use
in simulated earthquake tests on reinforced concrete shear wall systems(2*4g). Similar
I data for frame-wall systems is part of a later research phase.
The recommendations developed by the research group, and detailed in Ref-
I erence 2.49, resulted from the investigation of over 300 dynamic analyses conducted on
lo-, 20-, 30- and 40-story shear wall computer models. Ten different input ground mo-
I tions were used of which six were 10 seconds in duration and four were 20 seconds in
duration. The assumed plastic hinge regions were typical of cast-in-place systems as

I were the moment-rotation relationship.


Since the plastic hinge zones and the moment-rotation relationships of pre-
cast systems may be different, the results of this investigation cannot be directly applied
I to precast systems. However the summary listed in Reference 2.49 is given below since
it illustrated the type of guidelines required for precast construction.
I The summary contains the following six points:
“1. An estimate of the maximum amplitude of rotational deformation that
I can be expected for a particular earthquake intensity and combination
of structure period and yield level may be obtained from charts such as

I 2.
shown in Fig. 2-42.
Maximum shears corresponding to a particular combination of earth-
quake intensity, structure period and yield level may be estimated using
I charts such as shown in Fig. 2-43.
Dynamic analysis indicates that the critical shears shown in Fig.
I 2-43 are often produced by ground motions that are not the same as
those producing the critical ductility demand. Consequently, a conser-
I vatism beyond that associated with using the largest value for each
response quantity is obtained by assuming that they occur simul-

I taneously.

2.55
I
I
I
I
I
I
I
I
pip. !A-42 - Rotatlofml ductility vs. fundamental period
I
I
I
I
I
Fig. 2-43 -Critical
FuND*MENT*L PEWOD. T. SEC

shear at base vs. fundamental period I


I
I
I
I
pig. 3-4 - Representative deformation history for hinging region of isolated walls
I
2.56
I
I
I 3. The maximum number of fully-reversed cycles of large amplitude that
can be reasonably expected for strong ground motions of 20-second
I duration is six. The maximum total number of inelastic cycles, of small
and large amplitude, for the same duration of strong ground motion, is
I fully-reversed cycles of large amplitude. ‘lhe corresponding number of
inelastic cycles was eight.

I 4. Results of analyses indicate


should have the first maximum amplitude
that a representative loading program
of deformation early in the

I 5.
test with only one small inelastic cycle preceding it.
A suggested loading program is shown in Fig. 2-44. The value of maxi-
mum rotation in the hinging region, enex , and maximum shears corre-
I sponding to a particular earthquake intensity and combination of struc-
ture period and yield level can be obtained from charts such as shown in
I Figs. 2-42 and 2-43 respectively.
A specimen that sustains the applied loading without significant

I lossof strength can be considered adequate with respect to strength and


deformation capacity for the particular values of the significant design

I variables represented
earthquake intensity,
by the loading program.
structure
These variables include
period and yield level. Survival under
additional applications of the same loading history provides a measure
I of reserve capacity.
Satisfactory performance or survival in this context may be

I defined in terms of a specimen or structure


deformations and accompanying forces with
sustaining the imposed
a loss of strength not

I 6.
exceeding say, XI%, of it original strength.
A loading characterized by moments, shears and rotations all occurring
in phase, such as is commonly used in laboratory tests under slowly
I reversed loading, differs from dynamic inelastic reponse. Under dyna-
mic conditions, the shears are more sensitive to the higher modes of
I response. Consequently, they change direction much more rapidly than
do moment and rotation. For this reason, the commonly used test
I method of applying forces and deformations in phase represents a more
severe loading condition when compared to typical dynamic response.”

I 2.8 Performance of Precast Systems in Previous Earthquakes

One of the major contributions which enters into the development of an


I
2.57
I
I
earthquake-resistant design is the response and performance of structural systems during I
previous earthquakes. Many of the present building code provisions were developed from
observation and evaluation of earthquake damage. This is especially true for reinforced I
concrete structures where evaluation of earthquake damage has provided provisions for
ductility, reinforcement anchorage, confinement steel and static design forces.
Information concerning the seismic performance of precast concrete struc-
I
tures has been limited by the small number of precast structures in areas of high seismic
activity. However, the recent development of the precast industry in seismic regions
I
should provide a future basis for obtaining additional data.
When future seismic activity occurs, it is essential that precast performance I
is properly evaluated, recorded and published for undamaged as well as damaged struc-
tures. In the past the majority of earthquake reports concentrated on damaged struc- I
tures. Little information has been reported on the many precast structures which have
withstood major seismic accelerations. ‘lhus the engineer has a feel for what will not
work, but no confidence in what will work. With the variety of available precast details
I
any information on good or bad performance would be beneficial.
The following data has been recorded on the performance of precast struc-
I
tures during major earthquakes.
The Prince William Sound, Alaska, Earthquake I
Date: March 27, 1964
Magnitude: 8.4 (50 after shocks with magnitudes between 6 and 6.7 were recorded I
for three days after the major shock)
At the time of the earthquake twenty-seven low rise structures had been built
in the Anchorage area using precast, prestressed concrete members. Most of the struc-
I
tures contained precast prestressed single T-beam roof and floor systems supported on
precast or masonry bearing walls.
I
Five of the twenty-seven structures utilizing precast concrete collapsed
during the duration of the earthquake. Two of the structures were under construction at I
the time while the other three had insufficient lateral stability for the forces
developed. The remaining twenty-two structures suffered little or no damage(2.50). I
Damage to the precast systems occurred mainly at the connections of precast
members. The precast members suffered little or no distress. The most common failure
occurred at the diaphragm boundary connection which consisted of reinforcing bars
I
embedded in l-1/2 in. thick T-beam flanges, as shown in Fig. 2-33(b). Failure occurred
by pull out of a bar from one of the flanges. Inspection of the specimen revealed that
I
the bars were initially placed near the top of the flange with very little cover. This
I
2.58
I
I
I emphasizes the importance of adequate coverage and embedment for proper connection
behavior.
I The largest precast diaphragm to withstand the earthquake without damage
had an area of 6,500 square feet (2.51). Larger diaphragms all suffered connection dam-
I age in one form or another.
Another common area of distress occurred at the connection of exterior

I precast panels to the main frame or diaphragm system. Welded embedded anchors show-
ed little ductility and failure often occurred in a brittle manner by pull out or weld

I fracture.
Other troublesome areas included lap splices, development of reinforcement
and unexpected torsional stresses. It was also speculated that shrinkage and temperature
I effects may have overstressed some connections before the earthquake occurred (2.51).
San Fernando, California, Earthquake
I Date:
Magnitude:
February 9, 1971
6.6 (200 aftershocks with a magnitude greater than 3.0 were recorded

I through March 1, 1971. One aftershock of magnitude 5.1 occurred


immediately after the main shock.)

I The most common recorded precast failure consisted of exterior precast


panels pulling away from roof or floor supports(2.52). Connection pullout usually ini-
tiated the failure and in many instances the connection consisted of drill-in anchors. The
I general performance of drill-in anchors was rather poor during the earthquake. The
anchors showed little signs of ductility and their small embedment length (usually less
I than six inches) makes them very Inefficient for earthquake structures.
Damage from the earthquake also suggested that more consideration should

I be given to the dynamic characteristics of structures, especially buildings with irregular


shapes and with variations in stiffness between floors (2.52).
Rumanian Earthquake
I Date: March 4, 1977
Magnitude: 7.2
I The Rumanian Rarthquake was the first earthquake to subject a large number
and variety of precast structures to large seismic accelerations. The excellent perfor-
I mance of these structures illustrated the earthquake-resistant capacity of precast con-
struction as well as setting a precedent for further development of the precast industry.

I The largest fully precast structure built at the time of the earthquake was a
nine story bearing wall apartment building. The earthquake design of this and other
precast structures was developed through forced vibration testing of quarter-scale ten
I
I 2.59
I
story structures. The test specimens were typical of bearing wall construction with
I
horizontal and vertical connections composed of welded reinforcing bars or interlacing of
protruding reinforcing hoops(2’54). I
Finte1(2’54) suggests that the response of the large panel precast structures
was within their elastic range. If the structures were designed to perform elastically, I
there would be no signs of ductile behavior. The minimal amount of damage and repair
after the earthquake suggests that elastic response is an efficient way to design precast
structures.
I
2.9 Structural Integrity I
2.9.1 Definition
Structural integrity can be defined as that quality of a structure which allows I
stresses to be transferred from one component of a structure to another component. In
the event of the failure or removal of a structural component, the stresses carried by I
that component have to be redistributed away from the damaged area. This ‘alternate
path” philosophy of design requires that a structure be tied together to resist forces not
usually encountered under normal service loads.
I
The purpose of designing for structural integrity is to prevent a local failure
from producing a catastrophic collapse. Lack of structural integrity can be compared to
I
a row of dominoes or a house of cards, where the loss of one element precipitates the
collapse of the remaining elements. Providing an alternate path for the loads through I
the use of structural continuity will serve to bridge the local element failure, preventing
a collapse that would be entirely disproportionate to the initiating cause (known as
“progressive collapse?.
I
This philosophy of damage confinement can be illustrated in the design of
other types of structures. Bulkheads seal off compartments on ocean-going vessels to
I
prevent a single break below the water line from sinking the entire ship and party walLs
in apartment complexes localize damage due to fire or smoke.
I
Until recently most codes have not addressed the concept of adequate struc-
tural integrity directly. The reason is that most buildings contain a certain amount of I
“integrity” built-in.
Cast-in-place, reinforced concrete buildings contain numerous continuity I
enhancing features. These include minimum temperature reinforcement, 1% minimum
column reinforcement, #5 bars at openings, #3 bars at 12” to anchor floors to walls, wall
reinforcement in both directions, etc. These features enable reinforced concrete struc-
I
tures to withstand nearly all abnormal loadings without sustaining a total collapse.
I
2.60
I
I Such arbitrary, continuity-enhancing criteria also exist in structural steel
buildings. These factors include minimum force connection design, minimum weld size
I and length, slenderness limits, residual moment capacity in framed or seated connections
and other features of normal structural steel frame design.
I Non-structural components also play a major role in preventing a progressive
collapse. Such elements as concrete block partitions, window mullions, non-composite
I floor toppings, chimneys and fascia panels all contribute to post-abnormal loading stiff-
ness.

I Precast, prestressed concrete structures, however, do not yet have such a


stock of continuity-producing details. A partial reason for this Is that precast concrete
construction Is a relatively new technique, compared to steel or cast-inplace concrete
I design. The post-abnormal loading field observations of precast structure behavior are
not as numerous as those of steel and concrete (which have weathered two world wars
I together). This gap is being filled by large-scale tests on mock-up structures (2.69, s
well as sophisticated numerical analyses of cyclic or abnormal behavior of prestressed

I Also crucial to prestressed structure connection detailing is the desire to

I maintain simplicity in details and erection procedure. The key advantage of precast
construction is its capability for a simple and speedy erection, with most of the compli-
cated details accomplished in the prestressing plant. As a consequence, the design
I efficiency brought about by this “building block” approach tends to eliminate the design
redundancies present in steel or concrete construction, which also eliminates many of the
I alternate load paths desired for abnormal loading response. Therefore, current research
is geared toward restoring some of these redundant load paths, thereby increasing struc-

I tural integrity, while also attempting to retain the simplicity of precast, prestressed
concrete connection details.

I 2.9.2 Philosophy of Design for Structural Integrity


In any design for structural integrity, the designer’s hope is that these provi-
I sions will never have to be used. It is hoped that normal, ultimate load factors will
account for any abnormal loading situations. ‘lhese provisions are only useful in that rare

I instance when loads exceed ultimate to the point that a crucial support member, usually
a wall or a column, has yielded to the point where it can no longer transfer its assigned
loads effectively. These loads must then seek an alternate path for transfer to the
I ground.

I
2.61
I
I
I
I
I
I
(a) Immediate local damage (b) Progressive collapse
I
Fig. 2-45 - PaiIure modes of Roman Point collapse (Ref. 2.80)
I
The need for continuity of this type was demonstrated tragically in 1968 when
the 24 story Ronan Point apartment tower in London, England partially
explosion on the 18th floor caused an exterior
collapsed.
wall panel to be blown out, initiating
A gas
a
I
progressive collapse upward to the roof and then downward almost to the ground, due to
debris loading of the floors (See Fig. Z-45). As such, the initial damage was relatively
I
minor, but the resulting collapse was 23 times as damaging, demonstrating the structure%
inability to bridge over the local failure, i.e., its lack of integrity. I
Design forces at Ronan Point were resisted solely by the effects of bearing,
friction and bond, yet the building design fully complied with the then applicable building I
codes. After loss of the 18th floor wall panel, the structure was subjected to loads which
were unanticipated and, consequently, undesigned for, causing the collapse.
The Ronan Point disaster caused engineers and others to recognize that other
I
“abnormal”, unpredictable loadings can and sometimes do occur. This may include such
things as vehicle or airplane crashes, fires, floods, landslides, etc. In areas of low seis-
I
mic probability, earthquakes could be considered “abnormal loads”. This presents some-
thing of a dilemma: How does one design for loads which are, by nature, totally unpre- I
dictable? ‘lhe first approach, that of strengthening each element to resist any applied
abnormal load, is clearly uneconomical, as well as not totally attainable, given the un- I
predictable nature of these loads. A more logical approach would be to allow removal of
any structural element, at random, and designing the structure to transfer loads to the
other elements accordingly. ‘Ibis approach is still unrealistic in the sense that rarely
I
does an element simply Waporixe”
a yielded condition.
from the structure. It usually remains in place, but in
I
I
2.62
I
I
I Therefore, the most reasonable design approach would be to designate which
members are “allowed” to yield and which members are to be Wrong elements” (meant
I to stand firm at all cost). The member to be yielded is assumed to fail in a ductile
manner, allowing a semi-static type of load transfer. The post-yielded member also
I retains some of its original height, thereby limiting the deflection of the elements
above. This design approach is presented more completely in Reference 2.69.
I 2.9.3 Application to Severe Earthquake Resistance
Design for structural integrity is not meant to be a substitute for earthquake
I design. On the contrary, structural integrity provisions (as currently proposed) are bare
minimum requirements for 5 structure and any earthquake resistant design require-
I ments are in addition to the minimum integrity provisions.
This does not mean that design for integrity and continuity will not help in an

I earthquake situation. The ties used to hold elements together under abnormal loads will
also provide added ductility in the event of an earthquake. They are designed to sustain
large deflections without failure, which is precisely what is required to dampen severe
I earthquake vibrations in a prestressed building.
If the normal seismic provisions for a certain structure are exceeded and a
I member fails totally, structural integrity provisions will provide a secondary line of
defense. This is accomplished by shifting loads to other members, hopefully until the
I vibrations cease, allowing occupants time to evacuate the building.
2.9.4 Relationship to Service Load Design Procedures
I In general, design provisions for abnormal loads should be separated in func-
tion from the service load requirements for precast, prestressed construction.
I These load transfers should, in fact, be avoided under service conditions. The
additional member restraint added by these tie forces could cause unanticipated stresses

I due to creep, shrinkage, or temperature effects. Prestressed members could sustain


flexural cracking due to lack of negative moment resistance as the continuity ties re-
strain member ends.
I Reference 2.69 recommends using unstressed prestressing strand to tie the
structure together. These strands are capable of sustaining high strains without ruptur-
I ing and pull out tests confirm adequate ductility under abnormal loads. The large strain
capability of these strands also prevent them from transferring much load under service
I conditions. In other words, the strands would be passive under service loads and only
become active in the event of an abnormal loading situation, such as an earthquake or

I explosion.

I 2.63
I
2.9.5 Current Design Provisions for Structural Integrity I
1. AC1 318-77
The American Concrete Institute Building Code for Reinforced I
Concrete(2.25) does not address the problem of structural integrity
directly, but does provide certain requirements for precast, prestressed I
construction in general:
14.2.9 Walls shall be anchored to floors, roofs, or to columns,
pilasters, buttresses and intersecting walls.
I
16.2.2 In precast construction that does not behave monolithically,
effects at all interconnected and adjoining details shall be
I
considered to assure proper performance of the structural
system. I
16.2.3 Effects of initial and long-time deflections shall be con-
sidered, including effects on interconnected elements.
Design of joints and bearings shall include effects of all
I
16.2.4
forces to be transmitted, including shrinkage, creep, temper-
ature, elastic deformation, wind and earthquake.
I
2. U. S, Department of Housing and Urban Development - “Minimum
Property Standards for Multi-family Housh@’
I
Recently, the Prestressed Concrete Institute has proposed addi-
tions to the above document regarding structural integrity. These I
additions listed below have been accepted in principle by HUD.
a. In Section 601-18: “All buildings shall maintain a level of strength
that will minimize the possibility of local damage being propagat-
I
b.
ed beyond the immediate vicinity of such damage.”
In Section 603-10: “The following provisions for bearing wall
I
construction will satisfy the requirements of Section 601-18:
. ..Unless analysis indicates that a greater amount of ties are I
required, the following minimum ties shall be provided:
(1) Reinforcement in the direction of the floor span, connecting I
floor or roof units which abut over internal bearing walls, or
connecting units to end bearing walls, shall be provided at a
spacing not to exceed 8 feet. Reinforcement shall be capable
I
of developing the design (factored) force required, but not
leas than 1500 pounds per linear foot of distance measured
I
parallel to the support wall. Reinforcement may be connect-
I
2.64
I
I
I ed by mechanical methods, or be of sufficient length to

I transfer the specified force by bond to adjacent floor con-


struction. Unstressed strands or reinforcing bars, grouted in
joints between units, may be used for this purpose. Rein-
I forcement shall be discontinued at expansion joints.
(2) Reinforcement transverse to the direction of the floor span
I may be required to resist lateral load, differential settlement
and volume change. If analysis indicates tensile forces, these
I shall be resisted by reinforcement in or over transverse
walls. Such reinforcement may be continuous or may transfer

I load by bond or mechanical methods.


(3) Perimeter reinforcement shall be provided in or close to the
plane of the floor and roof slab, and within a distance of 48
I inches of the perimeter, capable of developing design (factor-
ed) force required, but not leas than 16,000 pounds. Rein-
I forcement may be continuous, or of sufficient length to
transfer force by bond. Alternatively, mechanical connec-

I tions may be used. A continuous reinforced bond beam may


be used as perimeter reinforcement in masonry bearing wall
construction.
I 3. PC1 Committee on Precast Concrete Bearing Wall Buildings(2*71)
This PC1 Committee report suggests a number of design criteria
I for resisting abnormal loads in precast concrete bearing wall buildings.
Some of these guidelines are loosely based on the British Unified Code
I (put into effect after the Ronan Point collapse) while other guidelines
are formulated from recent test data.

I Fig. 2-46 shows the recommended nominal tie placement within


the structure, with minimum design forces in these nominal ties equal
to:
I Tl 1500 (lb/ft) x span of floor slabs (ft), (cross tie)
T2 16,000 lb, peripheral tie
I T3 2-l/2% of service load on wall
1500 (lb/ft) x distance between ties (ft), (longitudinal tie)
I T4 3000 (lb/ft) x distance between ties (ft), (vertical tie)
These forces are required for considerations of structural inte-

I grity if the design does not require cross ties or continuity reinforce-

I 2.65
I
I
I
I
I
0 Recommended tie forces b) lppioal tie errangement
I
pig. 2-46 - Recommendations for nominal tie placement (Ref. 2.71)
ment of greater strength. I
Another requirement is that the structure withstand a lateral
design load equal to 2% of the building’s vertical service load. I
It is clear that tie requirements for earthquake resistance would
be more stringent than those presented above, but the locations of tie
placement would be similar.
I
Portland Cement Association Tests(2.69)
4.
The Portland Cement Association has conducted precast large
I
panel general structural integrity tests for the Department of Housing
and Urban Development and has presented methods for the design of I
structural ties to resist abnormal loads. The notation is the same as
that used by PC1 CT1through T4), so one can refer to Fig. 2-4’7 for tie
locations.
I
Fig. 2-47 presents a method of calculating the required capacity
of the transverse ties (Tl). Peripheral ties (T2) are to be designed as
I
transverse (T1) or longitudinal (T2) ties, depending on their orientation
(compared to the 16 kip arbitrary force for T2 recommended by PC& I
Longitudinal ties (T2) can be designed based upon the chart and table
presented in Fig. 2-48. Unstressed prestressing strand is recommended I
for Tq to prevent tie rupture and promote floor suspension action (See
Fig. 2-49). Vertical ties (T4) may be designed according to the chart in
Fig. 2-50 for horizontal shear. Tension force due to the possible loss of
I
a lower panel and subsequent suspension of upper panels should also be
accounted for in the final design of tie T4.
I
I
2.66
I
I
I
I
I
I
I
I
Fig. 2-47 - Design chart for transve.rse force at any story (T1) (Ref. 2.69)

I
I
I
I
I
I
I
I
I
I F&!&46-Deaignehartfor interior langitudinal ties (Ref. 2.69)

I
I 2.67
I
I
I
-Wall panels

I
I
I
Fig. 2-49 - suspension mechanism by means of longitudinal ties
I
I
I
I
I
I
I
I
N-IOVERTICAL CONNECTION / TIE
SHEAR FORCE “V,.” AT EACH STORY
I
C-I FOR ANY WiLL HEIGHT _

I
I
Fig. P-60 - Des&p chart for horizontal shear force at any story Wet. 2.69)

I
2.68
I
I
I 2.10 Hurricanes, Tornadoes and Other Wind Loadings

I 2.10.1 -Comparison with Earthquake Loading


The design of a typical structure for wind load is very similar to an earth-

I quake load design. In both cases, dynamic loads are approximated by equivalent static
loads applied laterally to the structure. For earthquakes, this static load is a small
percentage of the actual earthquake load, while static wind loads are representative of
I the actual maximum wind force.
The source of these loads differ, however. Earthquake loads are generated by
I the mass of the building itself, larger masses creating larger lateral forces. Wind forces
are externally applied. Their magnitude is dependent on local climate, building shape,
I orientation, and surface texture. Therefore, while excessive mass is a detriment to the
earthquake resistance of a structure, it is an asset to the structure’s wind resistance.

I While earthquake design is concerned with ultimate performance in the


inelastic range, normal wind design is usually governed by serviceability requirements.
Maximum drift coefficients are formulated from human perception of movement, or
I crack control criteria and are usually more stringent than strength criteria for mid-rise
and high-rise structures. Another concern in a design for wind loads is that of uplift
I pressures on roofs and eaves, as well as suction pressures on leeward walls. These unique
design criteria should make it clear that a design to resist earthquakes (though loads are
I more severe) will not necessarily resist wind loads, or vice-versa.
2.10.2 Building Code Requirements

I ANSI A58.1- 1980 (DRAFT)


‘Die 1980 ANSI Standard revision differs from the previous 1972 Standard in

I that the allowable wind pressure tables have been replaced by pressure formulas and
coefficients. ‘this was done to reflect the proliferation of hand calculators which make
calculations easier than table reading.
I Design wind pressures are a function of velocity pressure, q, a gust factor, C,
and a pressure coefficient, C. Velocity pressure, q, is also a function of exposure* Rx,
I Importance, I, Hurricane vulnerability, H, and air velocity, V. A procedure for estimat-
ing the dynamic effect of wind gusts is suggested in the document. Pressure on building
I components and cladding are given special attention, since these pressures can be much
higher than the average pressure on the building. Tornadoes are not considered, but

I references are suggested which deal with design for tornado forces. Wind tunnel tests
are recommended for unusual conditions or in lieu of the above analytical procedure.

I
2.69
I
I
BOCA BASIC BUILDING CODE/l978 I
The BOCA Code follows ANSI A58.1-1972 recommendations closely. How-
ever, BOCA’s arbitrary assumption of cladding pressure equal to l-1/2 times the overaU I
pressure is unconservative at building corners, where the component pressure can be as
high as three times the overall pressure. It is expected that this assumption will be
revised to follow the ANSI 1980 revision recommendations.
I
UNIFORM BUILDING CODE, 1979 EDITION
The Uniform Code only specifies wind pressures as lateral forces applied to
I
the vertical projection of a building. No mention is made of cladding pressures or wall
suction pressures. These deficiencies are partially compensated for by specifying lateral I
pressures which are much higher than ANSI recommendations. This approach is unduly
conservative for structural design, yet possibly unconservative for cladding at building I
corners, windows and for negative pressure resistance.
2.10.3 Common Practice I
As was shown in the previous section, most building codes address conven-
tional wind storms and hurricanes only. A design for these loads is concerned mainly I
with service load conditions such as lateral drift control and movement perception.
Ultimate strength becomes critical only for low-rise buildings or for suction pressures on I
ClEWMing. An aeroelastic analysis is required for structure types that tend to vibrate
under the dynamic action of wind gusts. Precast concrete structures fare better than
steel in this respect, since greater mass means less acceleration.
I
Hurricane loads are anticipated by designing for conventional maximum 58 or
199 year recurrence wind speeds and adding a hurricane factor, H, depending on the
I
proximity to the coastline (ANSI A58.1-1980). As in earthquake resistant design, it is
expected that a severe hurricane could stress the structure to near ultimate strength and I
also cause some non-structural damage. This can be tolerated since such storms are
extremely rare and to design for such a rare occurrence would be economically unjustifi-
able. Structural ductility therefore becomes necessary to prevent a brittle failure from
I
occurring during the hurricane passage.
Wind-borne objects become missiles that can penetrate a building’s skin.
I
Precast wall panels provide an effective barrier against these missiles, though large glass
openings could subvert this advantage. .
I
Design for tornado resistance is normally not done for conventional struc-
tures. The reason for this is that although tornadoes are the severest of the wind storm I
types, they are also small in scope, leaving damage paths normally less than a few hun-
I
2.70
I
I
I dred yards wide. This makes it impossible to predict where they will occur. Typical
recurrence rates for tornadoes range in the thousands of years. Only for critical installa-
I tions such as nuclear power plants are such remote risks given any consideration. How-
ever, recent studies of tornado paths reveal lower wind speeds (100 to 250 mph) and
I pressures than previously expected, bringing tornado resistant design into the realm of
plausibility.
I 2.10.4 Recommendations
For conventional winds, Reference 2.73 suggests limiting drift to h/1000 for
I prefabricated panel buildings. This value lies between the AC1 maximum of h/500 and
the CEB recommendations of h/2000 for the limit state of cracking.
I Precast concrete shear walls can be used to effectively reduce drift in high-
rise steel structures(2*74). These panels could also serve to limit non-structural damage
I caused by a wildly gyrating ductile steel frame in the event of an earthquake.
Partial fixity at wall-floor panel joints under service wind loads is recom-

I mended by Reference 2.75 as a means of limiting lateral deflections. Fig. 2-51 shows
how the lateral drift varies, depending on the fixity of the wall-floor connection. A 20%
fixity serves to cut the total drift in half for an average mid-rise precast large-panel
I building. If the joint is not reinforced, however, it could fracture under earthquake loads
and the joint fixity benefits would be lost. Building codes utilizing an equivalent static
I earthquake load tend to obscure this point.

I
I
I
I
I
I DEGREibF OFW-ALL -FiOOR
CONNECTION

Fig. Z-51- Relative displacement vs. fixity of wall-floor connecti&~‘s)


I
I 2.71
I
Examination of buildings after a severe hurricane reveal that structural
I
damage is not as pronounced as the damage to cladding and other secondary structural
components(2*76). In many cases, these secondary failures led to a progressive failure of I
the rest of the structure.
A case in point is the Central Fire Station at Lubbock, Texas during the I
Lubbock storm of May 11, 1970. Though this storm more closely resembled a tornado, its
Will serve to illustrate how a wind-induced progressive collapse can occur. Failure of
this structure was attributed to the failure of the two overhead doors facing South.
I
South winds caused the doors to buckle inward, allowing wind to fill the structure.
Internal pressure and external suction caused the structural clay tile east wall to fail
I
outward. ‘lhe steel joist roof supported by this wall then collapsed on top of the fire-
fighting equipment, preventing their use. The 1980 ANSI Standard adds an importance I
coefficient, I, to the basic wind speed estimate to help prevent damage such as this to
essential facilities. I
In precast panel construction, joint continuity under severe wind load is
recommended. This may prevent a progressive collapse from occurring due to loss of a
wall panel from suction pressure. Experience has shown such a panel loss to be rare,
I
because of the mass-stability of precast bearing wall construction, but the possibility
still exists. Section 2.9 discusses design for structural integrity in greater detail.
I
It was shown that doors and large windows possessthe greatest vulnerability
in a hurricane or tornado strike. In addition, airborne debris act as missiles which can I
penetrate the building at these weak points. Unless these structural openings are uni-
formly distributed around the building, broken windows on the windward side can cause a I
failure as described above (the Lubbock Fire Station). Provision should be made for
venting these internal pressures during a hurricane or tornado strike.
Roof connections also deserve special attention since they must resist roof
I
and eave uplift forces. Again, if these joints are provided with a nominal continuity, as
recommended in Sect. 2.9, they should be able to resist these uplift forces.
I
Joint fatigue can be a problem for light structural systems that are prone to
wind flutter. The rigidity of precast concrete systems is usually adequate to prevent I
these vibrations from occurring, although fatigue due to wind gusts can still occur.
As mentioned earlier, tornado wind speeds are currently thought to be lower I
than estimated previously. These revised wind speedsare based mainly on the analysis of
tornado damage to %learP structures. That is, to structures of sufficient aerodynamic
simplicity to allow reasonably accurate calculations of the wind speed required to precip-
I
itate failure.
I
2.72
I
I
I
I
I
I
I
I
I
I
Fig. 2-52 - Combined extreme winds distribution (1 mph = 1.6 km/hr)fRef. 2.77)
I
Several methods have been proposed to assess the probability of a damaging
I tor,,ado(2’77, 2*78)m Fig. 2-52 shows a design wind chart derived from an analysis by Wen
and Chu(2*‘7). lhese mathematical probability analyses necessarily rely somewhat on

I subjective assumptions and the results too often indicate a degree of precision that has
very little meaning.
What is more relevant is the increased cost an owner is willing to absorb to
I provide some degree of tornado resistance and to what extent building codes should
require consideration of tornado velocity winds.
I Most precast, prestressed concrete buildings would require very little modifi-
cation to withstand exceptionally high wind velocities (2.79), particularly if ultimate
I strength of the components and connections is considered, as is the case with earthquake
resistant design.

I
I
I
2.73
I
I
BRFERBNCES - PART 2 I
2.1 Derecho, A. T. and Fintel, M. “Chapter 12: Earthquake Resistant Structures”,
Handbook of Concrete Engineering (Fintel, M. ed.), Van Nostrand Reinhold Com-
I
pany, New York, 1974.
2.2 Bolt, B. A., “Chapter 2: Causes of Earthquakes”, Earthquake Engineering
I
(Wiegel, R. ed.), Prentice-Hall, Inc., Fnglewood Cliffs, NJ, 1970.
2.3 Newmark, N. M., “Chapter 16: Current Trends in the Seismic Analysis and Design I
of High-Rise Structures ‘I, Earthquake Engineering (Wiegel, R. ed.), Prentice-Hall,

2.4
Inc., Englewood Cliffs, NJ, 1970.
Clough, R. W., “Chapter 12: Earthquake Response of Structures”, Earthquake
I
Engineering (Wiegel, R. ed.), Prentice-Hall, Inc., Englewood, NJ, 1970.

2.5 Llorente, C., Becker, J. M. and Roesset, J. M., “Effect of Nonlinear Inelastic
I
Connection Behavior on Precast Panelized Shear Walls”, Reinforced Concrete
Structures Subjected to Wind and Earthquake Forces, Publication SP-63, American
Concrete Institute, Detroit, 1980. I
2.6 Powell, G. and Schricker, V., “Ductility Demands of Joints in Large Panel Struc-
tures”, ASCE Fall Convention, San Francisco, October, 1977, Preprint 3022. I
2.7 Brankov, G. and Sachanski, S., “Response of Large Panel Buildings for Earthquake
Excitation in Non-elastic Range I’, Proceedings of the Sixth World Conference on
Earthquake Engineering, New Delhi, India, 1977. I
2.8 Newmark, N. M. and Hall, W. J., “Dynamic Behavior of Reinforced and Prestress-
ed Concrete Buildings under Horizontal Forces and the Design of Joints”, Preli-
minary Publication, Eighth Congress of International Association for Bridge and
I
Structural Engineering, New York, 1968, pp. 586-613.

2.Y Green, N. B., “Earthquake Resistant Building Design and Construction”,


I
Van Nostrand Reinhold Company, New York, 1978.

2.10 Parme, A. L., “American Practice in Seismic Design”, PC1 Journal, Vol. 17, No. 4, I
July/August, 1972, pp. 31-44.
2.11 Phillips, W. R. and Sheppard, D. A., “Plant Cast Precast and Prestressed Con-
crete”, the Prestressed Concrete Manufacturers Association of California, Inc.,
I
1980.

2.12 Dorwick, D. J., Earthquake Resistant Design, John Wiley d( Sons, New York, 1977.
I
2.13 Seismology Committee, SEAOC. “Recommended Lateral Force Requirements
Commentary”, Structural Engineers Association of California, 1974.
and
I
2.14 Steinbrugge, K. V., “Chapter 9: Earthquake Damage and Structural Performance
in the United States”, Earthquake Engineering (Wiegel, R. ed.), Prentice-Hall, Inc.,
Englewood Cliffs, NJ, 1970.
I
I
2.74
I
I
I 2.15 Uniform Building Code, International Conference of Building Officials,
Whittier, CA, 1979.
I 2.16 ATC-3 (Applied Technology Council), “Tentative Provisions for the Development
of Seismic Regulations for Buildings”, National Bureau of Standards, U. S. De-
partment of Commerce, Washington, DC, June, 1978.
I 2.17 Anonymous, “Tough Rules To Make Buildings Quake-resistant”, Business Week,
October 6, 1980, pp. 112E-F.
I 2.18 Park, R. and Paulay, T., Reinforced Concrete Structures, John Wiley & Sons,
New York, 1975, pp. 545-609.
I 2.19 Blume, J. A., Newmark, N. M. and Corning, L. H., Design of Multi-story Relnforc-
ed Concrete Buildings for Earthquake Motions, Portland Cement Association,
Chicago, IL, 1961.
I 2.20 Clough, R. W., ‘Effect of Stiffness Degradation on Earthquake Ductility Require-
ments”, Report No. 66-16, Structural Engineering Laboratory, Berkeley, Univer-
I 2.21
sity of California, October, 1966.
Fintel, M., “Quake Lessons from Managua: Revise Concrete Building Design?“,
Civil Engineering, Vol. 43, No. 8, August, 1973, pp. 60-63.
I 2.22 Paulay, T., “Earthquake-Resisting Shear Walls - New Zealand Design”, AC1 Jour-
nal, Vol. 77, No. 3, May-June, 1980, pp. 144-152.
I 2.23 Allen, C. M., Jaeger, L. G. and Fenton, V. C., “Ductility in Reinforced Concrete
Shear Walls”, AC1 Publication SP-36, Responseof Multi-story Concrete Structures
to Lateral Forces, American Concrete Institute, Detroit, MI, 1973, pp. 97-118.
I 2.24 Derecho, A. T., Ghosh, S. K., Iqbal, M. and Fintel, M., “Strength, Stiffness and
Ductility Required in Reinforced Concrete Structural Walls for Earthquake Resis-
I tance”, AC1 Journal, Vol. 76, No. 8, August, 1979, pp. 875-896.
2.25 AC1 Committee 318, “Building Code Requirements for Reinforced Concrete
I 2.26
(AC1 318-77)“, “American Concrete Institute, Detroit, MI.
Englekirk, R. E., “An Evaluation of the State of Art in the Design and Construc-
tion of Prefabricated Buildings in Seismically Active Areas of the United State”,
I Proceedings, Workshop on Earthquake Resistant Reinforced Concrete Building
Construction, University of California at Berkeley, July, 1977.

I 2.27 Arya, A. S., “Use of Prestressed Concrete in Earthquake Resistant Structures”,


Proceeding, Fourth Symposium on Earthquake Engineering, Roorkee, India,
Nov. 1970.
I 2.28 Popov, E. P., “Seismic Behavior of Structural Subassemblies”, Journal of the
Structural Division, ASCE, Vol. 106, No. ST7, July, 1980, pp. 1451-1474.

I 2.29 Anonymous, “Eccentric Bracing Is Key to Seismic Resistance”, Engineering News


Record, October 25, 1979, pp. 32-33.

I
2.75
I
I
2.30 Roeder, C. W. and Popov, E. P., “Eccentrically Braced Steel Frames for Earth- I
quakes”, Journal of the Structural Division, ASCE, Vol. 104, No. ST3, March, 1978,

2.31
pp. 391-412.
Scott, N. L., “Developments in Precast Framing ‘I, Concrete International, Ameri-
I
can Concrete Institute, Vol. 2, No. 11, November, 1980, pp. 60-66.
2.32 Becker, J. M. and Lorente, C., Seismic Design of Precast Concrete Panel Build-
I
ings”, Proceedings, Workshop on Earthquake Resistant Reinforced Concrete Build-

2.33
ing Construction, University of California at Berkeley, July, 1977.
Hawkins, N. M., “Stateof-the-Art Report on Seismic Resistance of Prestressed
I
and Precast Concrete Structures: Part 2 - Precast Concrete”, PC1 Journal,
Vol. 23, No. 1, Jan-Feb, 1978, pp. 40-58. I
2.34 Fiorato, A. E. and Corley, W. G., “Laboratory Tests on Earthquake Resistant
Structural Wall Systems and Elements”, Proceedings, Workshop on Earthquake
Resistant Reinforced Concrete Building Construction, University of California at
I
Berkeley, July, 1977.
2.35 Spencer, R. A. and Nellie, D. S., “Cyclic Tests of Welded Headed Stud Connec-
tions”, PC1 Journal, Vol. 21, No. 3, May-June, 1976, pp. 70-83.
I
2.36 PC1 Design Handbook, Second Edition,
Chicago, IL, 1978.
Prestressed Concrete Institute, I
2.37 Venuti, W. J., “Diaphragm Shear Connectors Between Flanges of Prestressed
Concrete T-Beams”, PC1 Journal, Vol. 15, No. 1, February, 1970, pp. 67-78. I
2.38 Shemie, M., “Bolted Connections in Large Panel System Buildings”, PC1 Journal,
Vol. 18, No. 1, Jan-Peb, 1973, pp. 27-33. I
2.39 Pillai, S. U. and Kirk, D. W., ‘Ductile Moment Resisting Beam-Column Connec-
tions on Precast Concrete ‘I, Paper presented at AC1 Fall Convention, Washington,
DC, Oct. 28-NOV.2,1979.
I
2.40 Hawkins, N. M. and Mitchell, D., “Historical Perspective”, Reinforced Concrete
Structures in Seismic Zones, Publication SP-53, American Concrete Institute, I
Detroit, 1977.
2.41 Park, R., “Design of Prestressed Concrete Structures”, Proceedings, Workshop on
Earthquake Resistant Reinforced Concrete Building Construction, University of
I
California, Berkeley, July, 1977.
2.42 Thompson, K. J. and Park, R., “Ductility of Prestressed and Partially Prestressed
I
Concrete Beam Sections”, PCI Journal, Vol. 25, No. 2, March/April, 1980,

2.43
pp. 46-70.
Thompson, K. J. and Park, R., “Cyclic Load Tests on Prestressed and Partially
I
Prestressed Beam-Column Joints”, PC1 Journal, Vol. 22, No. 5, Sept/Oct, 1977,
pp. 84-110. I
I
2.76
I
I
I 2.44 Lin, T. Y., Kuika, F. and Tai, J., “Design of Earthquake-Resistant, Prestressed
Concrete Structures”, Proceedings, Workshop on Earthquake Resistant Reinforced
I 2.45
Concrete Building Construction, University of California at Berkeley, July, 1977.
Code of Practice for General Structural Design and Design Loadings for Buildings
(NZS 4203: 1976), Standards Association of New Zealand, 1976.
I 2.46 Draft Code of Practice for the Design of Concrete Structures (DZ 3101), Stan-
dards Association of New Zealand, 19’78.
I 2.47 Dowrick, D. J., Kulka, F. and Park, R., “Connections Between Precast Prestressed
Concrete Members in Buildings”, FIP Commission on Seismic Structures, Interim
Report by Working Group, April, 1979.
I 2.48 Uzumeri, S. M., “Strength and Ductility of Cast-In-Place Beam-Column Joints”,
Reinforced Concrete Structures in Seismic Zones, Publication SP-53, American
I Concrete Institute, Detroit, 1977, pp. 293-350.
2.49 Derecho, A. T., Iqbal, M., Pintel, M. and Corley, W. G., “Loading History for Use
I in Quasi-Static Simulated Earthquake Loading Tests‘I, Reinforced Concrete Struc-
tures Subjected to Wind and Earthquake Forces, Publication SP-63, American
Concrete Institute, Detroit, 1980.

I 2.50 Kunze, W. E., Sbarounis, J. A. and Amrhein, J. E., “The March 27 Alaskan Earth-
quake-Effects on Structures in Anchorage”, AC1 Journal, Vol. 62, No. 6, June,
1965, pp. 635-648.
I 2.51 Wood, F. J. (Editor-in-Chief), The Prince William Sound, Alaska, Earthquake of
1964 and Aftershocks, U. S. Department of Commerce Environmental Science
I 2.52
Services Administration, Publication 10-3, Vol. 2, 1967.
Lew, H. S., Leyendecker, E. V. and Dikkers, R. D., Engineering Aspects of the
1971 San Fernando Earthquake, United States Department of Commerce, Building
I Research Division, Building Science Series 40, December, 1971.
2.53 Sharpe, R. L., ‘the Earthquake Problem”, Reinforced Concrete Structures in
I 2.54
Seismic Zones, Publication SP-53, American Concrete Institute, Detroit, 1977.
Fintel, M., “Performance of Precast Concrete Structures During Rumanian Earth-
quake of March 4, 1977”, PC1 Journal, Vol. 22, No. 2, March-April, 1977, pp. lo-
I 15.
2.55 Kunze, W. E., Sbarounis, J. A. and Amrhein, J. E., “Behavior of Prestressed Con-
I crete Structures During the Alaskan Earthquake”, PC1 Journal, Vol. 10, No. 2,
April, 1965, pp. 80-91.

I 2.56 Hanson, N. W., %eismic Resistance of Concrete Frames With Grade 60 Rein-
forcement”, Journal of the Structural Division, ASCE, Vol. 97, No. ST6, June,
1971, pp. 1685-1700.

I 2.57 CoIaco, J. P. and Siess, C. P., “Behavior of Splices in Beam-Column Connections,


“Journal of the Structural Division, ASCE, Vol. 93, No. ST5, October, 1967,
pp. 175-193.
I
I 2.77
I
2.58 Bresler, B. and Bertero, V. “Behavior of Reinforced Concrete Under Repeated I
Loading”, Journal of the Structural Division, ASCE, Vol. 94, No. ST6, June, 1968,

2.5Y
pp. 1567-1590.

Birkeland, P. W. and Birkeland, H. W., “Connections in Precast Concrete Con-


I
struction”, AC1 Journal, Vol. 63, No. 3, March, 1964, pp. 345-368.

2.60 Freeman, S. A., ‘Seismic Design Criteria for Multi-story Precast Prestressed
I
Buildings”, PC1 Journal, Vol. 24, No. 3, May/June, 1979, pp. 62-88.

2.61 Fintel, M., “Resistance to Earthquakes-Philosophy,


sponse of Multi-story
Ductility and Details”, Re-
Concrete Structures to Lateral Force, Publication SP-36,
I
2.62
American Concrete Institute, Detroit, 1973, pp. 75-95.

Lewicki, B and Pauw, A., “Joints, Precast Panel Buildings - State of the Art
I
Report No. 2”, Planning and Design of Tall Buildings, Vol. III, American Society of
Civil Engineers, New York, 1972, pp. 171-190. I
2.63 Zeck, U. I., “Joints in Large Panel Precast Concrete Structures”, R 76-16, De-

2.64
partment of Civil Engineering, Massachusetts Institute of Technology, Jan, 1976.

Frank, R. A., “Dynamic Modeling of Large Precast Panel Buildings Using Finite
I
Elements With Substructuring”, R76-36, Department of Civil Engineering, Massa-
chusetts Institute of Technology, August, 1976. I
2.65 Hassan, F. M. and Hawkins, N. M., “Anchorage of Reinforcing Bars for Seismic
Forcesl’, Reinforced Concrete Structures in Seismic Zones, Publication SP-53,
American Concrete Institute, Detroit, 1977, pp. 387-416. I
2.66 Hassan, F. M. and Hawkins, N. M., “Prediction of the Seismic Loading Anchorage
Characteristics of Reinforced Bars ‘I, Reinforced Concrete Structures In Seismic
Zones, Publication SP-53, American Concrete Institute, Detroit, 1977, pp. 417-
I
438.

2.67 Aswad, A., %elected Precast Connections: Low-Cycle Behavior and Strength”, I
Proceedings of the 2nd U. S National Conference on Earthquake Engineering,

2.68
Stanford University, California, August 22-24, 1979.

Maison, B. F. and Popov, E. P., %yclic Response Prediction for Braced Steel
I
Frames”, Journal of the Structural Division, ASCE, Vol. 106, No. ST7, July, 1980,
pp. 1401-1416. I
2.69 Schultz, D. M., Burnett, E. F. P. and Fintel, M., “A Design Approach to General
Structural Integrity”, Design and Construction of Large-Panel Concrete Strue-
tures, Office of Policy Development and Research, Department of Housing and
I
Urban Development, Washington, DC, October, 1977.

2.70 Thompson, K. J. and Park, R., ‘Seismic Response of Partially Prestressed Con-
crete”, Journal of the Structural Division, ASCE, Vol. 106, No. ST8, Proceedings
I
Paper 15598, August, 1980, pp. 1755-1775.
I
I
2.78
I
I
I 2.71 Speyer, I. J., “Considerations for the Design of Precast Concrete Bearing Wall
Buildings to Withstand Abnormal Loads”, for PC1 Committee on Precast Concrete
I 2.72
Bearing Wall Buildings, PC1 Journal, Vol. 21, No. 2, Mar-Apr, 1976, pp. 18-51.
Mast, R. F., ‘Seismic Design of 24-Story Building with Precast Elements”,
PC1 Journal, Vol. 17, No. 4, July-Aug, 1972, pp. 45-59.
I 2.73 Levy, Matthys P. and Varga, Istvan S., “High-Rise Panel Structures”, Journal of
the Structural Division, ASCE, New York, NY, Vol. 98, No. ST5, May, 1972,
I pp. 975-988.

2.74 Weidlingern, Paul, “Shear Field Panel Bracing I’, Journal of the Structural Division,

I 2.75
ASCE, New York, NY, Vol. 99, No. ST7, July, 1973, pp. 1615-1631.

Burnett, Eric F. P. and Rajendra, R. Clement S., “Influence of Joints in Panelized


Structural Systems”, Journal of the Structural Division, ASCE, New York, NY,
I Vol. 98, No. ST9, Sept, 1972, pp. 1943-1955.

2.76 Minor, Joseph E., Mehta, Kishor C. and McDonald, James R., “Failures of Struc-
I tures Due to Extreme Winds”, Journal of the Structural Division, ASCE, New
York, NY, Vol. 98, No. STll, Nov, 1972, pp. 2455-2471.
2.77 Wen, Yi-Kwei and Chu, Shih-Lung, “Tornado Risks and Design Wind Speeds”,
I Journal of the Structural Division,
Dee, 1973, pp. 2409-2421.
ASCE, New York, NY, Vol. 99, No. ST12,

I 2.78 Garsono, Robert C., Catalan, Jose Morla and Cornell, C. Allin, “Tornado Design
Winds Based on Risk”, Journal of the Structural Division, ASCE, Vol. 101, No. ST9,
Sept, 1975, pp. 1883-1897.
I 2.79 Mehta, Kishor C., “A Proposal for Tornado Resistant Design of Versa-Space Build-
ings”, McDonald, Mehta and Minor, Consulting Engineers, Lubbock, TX, NOV, 1975.

I 2.80 Fintel, M., Schultz, D. M., “A Philosophy for Structural Integrity


Buildings”, PC1 Journal, Vol. 21, No. 3, May-June 1977, pp. 46-49.
of Large Panel

I
I
I
I
I
I
I 2.79
I
I
I
I
I
I
I
I PART 3

I SELECTION AND DESIGN OF CONNECTIONS

I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
3. SELECTION AND DESIGN OF CONNECTIONS
I Satisfactory performance and economy of precast, prestressed concrete struc-
tures depends to a great extent on the proper selection and design of each connection.
I Several publications have addressed this subject (3.1-3.6). This section is based on infor-
mation from those and other publications, plus the experience of the study team. A
I great deal of input was received from designers and manufacturers
ed concrete, especially members of the Prestressed Concrete
of precast, prestress-
Institute Committee on

I Connection Details.

3.1 Criteria for Connections of Precast Concrete Members

I Precast concrete connections must meet a variety of design and performance


criteria, and not all connections are required to meet the same criteria. Some of the
I items discussed in this section are self-evident.
obvious and may require special consideration
Other requirements may not be so
or specification by the designer or owner

I of the structure.

3.1.1 Strength

I A connection must have the strength to resist the forces to which it will be
subjected during its lifetime. Some of these forces are apparent, caused by dead and live

I gravity loads, wind, earthquake,


and are frequently overlooked.
and soil or water pressure. Others are not so obvious
These are the forces caused by restraint of volume

I changes in the members and those required to maintain stability.


Volume changes are caused by temperature change, creep and shrinkage of the
concrete. Instability can be caused by eccentric loading (intentional or unintentional), as
I well as lateral loads from wind and earthquake. Very often, measures taken to resist
instability will aggravate the forces caused by volume changes, and vice-versa.
I ln addition to loads or forces that may be anticipated, the Engineer may choose
to provide additional capacity to resist some “unanticipated” or “abnormal” loads(3’7).

I These loads could include foundation settlement,


or others. Some loads, such as hurricane or earthquake
explosion, aircraft or vehicle collision
may be considered “normal” in

I one geographical area and “abnormal” in others.


These “abnormal” loads may be accommodated in connections by a capacity for
overload (which may be included in the standard load factors or ‘safety factors”), by
I
I 3.1
I
ductility within the connection, or by redundancy (alternate load paths) in the total
I
structure or within the connection.
The connection strength can be categorized by the types of stress that may be I
induced. These include:
a. Compressive (or bearing) I
b. Tensile
c.
d.
Flexural
Shear
I
e. Torsional
Many connections will have a high degree of resistance to one type of stress,
I
but little or no resistance to another. For example, a connection may have a high shear
capacity and little or no moment capacity. It may be unnecessary, or even undesirable to I
provide a high capability to resist certain types of stresses.
Methods of designing for the various strength requirements of connections are I
given in references 3.1 through 3.3. Most of these design methods employ fundamental
strength and analysis relationships. Some are based on empirical equations developed
from test data, or assumed behavior. Specific design aspects are discussed in more detail
I
in Section 3.3 and Part 4.
I
Friction
surface I
I
I
Fig. 3-l - Simple bearing connection
I
3.1.2 Ductility
I
Ductility is usually defined as the ability to accommodate relatively large
deformations without failure. In structural materials, ductility is usually measured by
I
the amount of deformation that occurs between first yield and ultimate failure. Apply-
ing this definition to connections creates some conceptual problems. An example, the
I
very simple bearing connection shown in Fig. 3-l.
I
3.2
I
I
I This connection, which would “yield” as soon as the friction was overcome,

I would not “fail” until it had deformed by a length b (assuming adequate bearing
strength). Under the above definition, it could, therefore, be considered “ductile”.
However, ductility in building frames is usually associated with moment resis-
I tance. This is particularly true in designing for earthquake forces, which is where con-
cerns over ductility are usually expressed. In concrete members with moment-resisting
I connections, the flexural tension is normally resisted by steel components, either rein-
forcing bars or structural steel members. First yield occurs in this steel component, and
I final failure may be from rupture of the steel, crushing of the concrete, or a failure of
the connection of the steel to the concrete.

I The above friction connection, while probably not reliable for use in earthquake
areas, may be satisfactory for some types of structures under some loading conditions,
such as wind. In that case the normal wind loading would be resisted by the friction. The
I “ductility”, assuming adequate bearing length, b, might be adequate for unusually strong
wind gusts, or other abnormal, very short duration loads.
I 3.1.3 Volume Change Accommodation
The combined shortening effects of creep, shrinkage and temperature drop can
I cause severe stresses on precast, prestressed concrete members and their supports, if the
end connections restrain movement. These stresses must be considered in the design, but
I it is usually far better if the connection will allow some movement to take place, thus
relieving the stresses.
I Section 4.4 of the PC1 Design Handbook(3’2) provides data and guidelines for
estimating the amount of volume change shortening that may take place.

I Most of the severe problems that have been caused by restraint of volume
change movements have appeared when relatively long members, usually stemmed deck
units, were welded to their supports at the bottom on both ends. When such members are
I connected only at the top, as shown in Fig. 3-2, experience has shown that VolUme
changes are adequately accommodated. On relatively short, heavily loaded members,
I such as beams, an unyielding top connection may attract negative moments if compres-
sion resistance is encountered at the bottom. This may be difficult to accommodate.
I Connections using cast-in-place concrete, similar to Fig. 3-3 have exhibited few
volume change problems. This is probably because micro-cracking and creep in the cast-

I in-place portion effectively relieves the restraint.

I
I 3.3
I
I
I
I
I
I
pig. 3-3 - Compoaite connection
Fig. 3-2 - Tppical double tee to beam connection
I
3.1.4 Durability
Evidence of poor durability is usually exhibited by corrosion
elements, or by cracking and spalling of concrete. Connections which will be exposed to
of exposed steel I
weather should have steel elements adequately covered with concrete, or should be
painted or galvanized. All exposed connections should be periodically inspected and
I
maintained.
Most precast, prestressed concrete is of high quality, and flexural cracking is
I
seldom a serious problem, provided tensile stresses are kept within code limits. How-
ever, local cracking or spalling can occur when improper details result ,in restraint of I
movement or stress concentrations.

3.1.4.1 Concrete encasement of exposed steel. Primarily because of appearance, the I


preferred method of protecting exposed steel connection elements is to cover them with
concrete, mortar or grout. Mix proportions, aggregate size and application procedures I
will vary with the size, location, and orientation of the element to be covered. For small
elements, such as recessed plates, it is often desireable to use a non-shrink, non-metallic
grout. Mixes containing chlorides should be avoided.
I
Patches in architectural panels and others that will be permanently
view will often require blending of white cement or color pigment to match the precast
exposed to
I
surface. Satisfactory results are usually dependent on experimentation and the experi-
ence of the workman. I
Anchoring the concrete or grout to the steel is a problem that is often over-
looked. Larger elements such as steel haunches can be wrapped with mesh or wire. For I
I
3.4
I
I
I
I (a) Heeded studs with wire tie
8

I E II

I (b) Notched rec&guler anchor with wire

I
I (c) Wiled refractory anchor

I Pig. 3-4 - Methods of anchoritg patches to recessed plates

I
recessed plates and similar elements, connections such as those shown in Fig. 3-4 can be
I used.

3.1.5 Fire Resistance


I Many precast concrete connections are not vulnerable to the effects of fire and
require no special treatment. For example, the bearings between slabs or stemmed units
I and beams do not generally require special fire protection. If the slabs or tees rest on
elastomeric pads or other combustible materials, protection of the pads is not generally
I needed because deterioration of the pads will not cause collapse. After a fire, the pads
could be replaced.

I Connections in which weakening by fire would jeopardize the structure’s


ity should be protected to the same degree as that required for the structural frame. For
stabil-

I example, an exposed steel bracket supporting a beam may be weakened enough by a fire
to cause the beam to collapse. Such a bracket should be protected. The amount of
protection depends on (a) the stress-strength ratio in the steel at the time of the fire and
I (b) the intensity and duration of the fire. Also many connections are simply stability
devices and are under little or no stress in service. A fire could substantially reduce the
I strength of such a connection but it would still perform effectively.

I
I 3.5
I
Connections which require a fire resistance rating will usually have exposed
I
steel elements encased in concrete
include enclosing with gypsum wallboard,
(See X1.4.1).
coating
Other methods of fire protection
with intumestic mastic, or spraying
I
with fire protection material.
There is evidence that exposed steel hardware used in connections is less sus- I
ceptible to fire-related strength reduction than other exposed steel members. This is
because the concrete elements provide a “heat sink”, which draws off the heat and re-
duces the temperature of the steel, Research on this subject Is recommended.
I
Additional
Reference 3.52.
information on fire protection of connections Is given in
I
3.1.6 Fabrication Simplicity I
Maximum economy of precast, prestressed concrete construction is achieved
when connection details are kept as simple as possible, consistent with adequate perfor-
mance and ease of erection. Furthermore, complex connections are more difficult to
I
control and will often result in poor fitting
erection and less satisfactory performance.
in the field. This can contribute to slow
I
Following is a list of items to consider in order to improve fabrication simplic-
ity. In many cases, some of these items must be compromised in order for the connec- I
tion to serve its intended function.
3.1.6.1 Avoid congestion. The area of the member in which the connection is made I
frequently requires large amounts of additional reinforcing steel, embedded plates,
inserts, blockouts, etc. It is not unusual for so many items to be concentrated into one I
area that there is very little room for concrete. It should be remembered that precast,
prestressed
top accessible.
concrete members are usually cast in permanent steel forms with only the
Placement of the reinforcement and hardware can often be likened to
I
“building a ship in a bottle”. If one item must be threaded in and around other items, for
example, labor costs can be significantly increased. Reinforcing bars and prestressing
I
strands, which appear as lines on drawings, have real cross-sectional dimensions (which
are larger than the nominal dimension because of the deformations); a fact to be consi- I
dered in the design phase. Reinforcing bar bends require minimum radii, which can cause
fit problems and leave some regions unreinforced.
to draw large scale details of the area in question.
If congestion is suspected, it is helpful I
In some cases, it may be economical to increase the member sizes just to avoid
congestion. Also, such details as dapped or recessed ends should be avoided unless neces-
I
sary. They require special reinforcement in a smaller space and are always congested.
I
3.6
I
I
I 3.1.6.2 Avoid penetration of forms. Since most precast, prestressed concrete members

I are cast in steel forms, projections


and costly to place.
which require cutting
Where possible, these projections
through the forms are difficult
should be limited to the top of the
member as cast. Even this inhibits finishing of the top surface especially on deck mem-
I bers or double tees and hollow-core slabs.

3.1.6.3 Minimize embedded items. Items which are embedded in the member, such as
I inserts, plates, reglets, etc., require plant labor to locate precisely and attach securely.
Therefore, these items should be kept. to a minimum.
I 3.1.6.4 Reduce post-stripping work. A plant casting operation is most efficient when
the product can be taken directly to the storage area immediately after it is stripped
I from the form. Any operations which are required after stripping and before placement
at the job site, such as special cleaning or finishing, or welding on projecting hardware,
I should be avoided, whenever possible. These operations require additional
work space, and added labor, often with skilled trades.
handling, extra

I The trade-off
times need to be evaluated.
between penetration of forms and post-stripping work will some-
An example of this is the embedded steel haunch shown in

I Fig. 3-5.
In Fig. 3-5(a), the operations that are required are:
(1) Cut hole in steel form with cutting torch.
I (2) Smooth rough cut edges.
(3) Place and secure steel haunch (requires a special device to suspend the
I (4)
steel and hold rigidly in place).
Caulk opening around the steel haunch.

I (5)
(6)
After all casts are made, patch the hole by welding.
Grind welded surface smooth.
In (b) the required operations are:
I (1) Weld plate to haunch at A in steel fabrication shop.
(2) Place and secure to form (easier than in (a) because it does not have to be
I suspended).
(3) After stripping the product, remove to special work area.
I (4)
(5)
Turn member on side to avoid overhead weld.
Weld on exterior portion at B.

I 3.1.6.5
(6) Remove member to storage.

Avoid non-standard tolerances. Dimensional tolerances which are specified to

I be more rigid than industry standards, are difficult to achieve. Connections which re-

I 3.1
I
I
Steel
ham-h
----
==Tzc=
I
Form
p.n.tmed I
fEk
(a) I
Fig. 3-5 - Alternate methods of casting column with
embedded steel haunch
I
quire close-fitting parts without provision for adjustment should be avoided. Standard
I
tolerances are given in PC1 and AC1 documents.

3.1.6.6 Use standard items. Wherever possible hardware items such as inserts, studs,
I
steel shapes, etc., should be standard items that are readily available, preferably from
more than one supplier. Custom fabricated or very specialized proprietary items add I
cost and may cause delays.
It also simplifies fabrication if similar items on a product or project are stan- I
dardized as to size and shape. For example, if half of the inserts are required to receive
3/4 in. diameter bolts, and the other half 1 in. diameter, use of all 1 in. bolts will be
simpler. There is also less chance of error. The same principle applies to reinforcing
I
bars, embedded plates, studs, etc.

3.1.6.7 Use repetitious details. It is very desirable to repeat details as much as pos-
I
sible. Similar details should be identical,
Once workmen are familiar with a detail,
even if it may result in a slight overdesign.
it is easier to repeat it than to learn a new
I
one. Also, it will require fewer form set-ups and improve scheduling.

3.1.6.8 l3e aware of material limitations. Examples of this are the radius requirements
I
for bending reinforcing

3.1.6.9 Allow alternates.


bars, standard lengths for certain sizes of inserts, etc.

Very often, a precast concrete manufacturer will prefer


I
certain details over others. The producer should be allowed to use alternative
or materials, provided the design requirements are met. Allowing alternate
methods
solutions will
I
often result in the most economical and best performing connection.
3.1.7 Erection Simplicity
I
Much of the advantage of precast, prestressed concrete construction is due to
I
3.8
I
I
I the possibility of rapid erection of the structure. To fully realize this benefit, and to
keep the costs within reasonable limits, field connections should be kept simple.
I In order to fulfill the design requirements, it is sometimes necessary to com-
promise fabrication and erection simplicity. Following is a list of items that should be
I considered during the selection, design and detailing of the connections that will facili-
tate speedy and safe erection.

I 3.1.7.1 Plan for the shortest possible hoist hook-up time. Hoisting the precast pieces is
usually the most expensive and time-critical process of the erection. Connections should
I be designed so that the unit can be lifted,
time.
set, and unhooked in the shortest possible
Before the hoist can be unhooked, the precast piece must be stable and in its final

I position.
inherently
Some shapes of precast units such as double tees and hollow-core
stable and require no additional connections before releasing
slabs are
the crane.

I Others, such as columns, deep beams, wall panels and single tees usually require some
supplemental shoring, guying, or fastening before the hoist can be unhooked. Preplanning
for the fewest, quickest, and safest possible operations that must be performed before
I releasing the hoist will greatly facilitate erection. Bearing pads, shims, or other devices
upon which the piece is to be set should be placed ahead of hoisting. Loose hardware
I that is requited for the connection should be immediately available for quick attachment.
In some cases, it may be necessary to provide temporary fasteners or leveling

I devices, with the permanent connection made after the hoist is released. For example, if
the permanent connection requires field welding, grouting, dry packing, or cast-in-place
concrete, erection bolts, pins, or shims can be used. These temporary devices must be
I given careful attention to assure that they will hold the piece in its proper position
during the placement of all pieces that are erected before the final connection is made.
I 3.1.7.2 Provide for adjustment. A certain amount of field adjustment at the connec-
tions is always necessary. Normal fabrication tolerances preclude the possibility of a
I perfect fit in the field. This is true not only when the precast pieces join to each other,
but, usually even more so, when the precast units must interface with other materials.
I Adjustment in the field is accomplished
holes for bolts and dowels, field welding, shims and grout.
through the use of slotted or oversize

I 3.1.7.3 Provide accessibility. Connections should be planned so that they are accessible
to the workman from a stable deck or platform. The type of equipment necessary to

I perform
sidered.
such operations as welding, post-tensioning, or pressure grouting should be con-
Operations which require working under a deck in an overhead position should be

I
I 3.9
I
avoided, especially for welding. Alternatives to anything that requires temporary scaf-
I
folding should be considered.Room to place wrenches on nuts and swing them in a large
arc should be provided for bolts. Dry-packing column or wall panel bases in a narrow I
excavation is difficult.

3.1.7.4 Standardize connection types. All connections which serve similar functions I
within the building should be standardized as much as possible. As workmen become
familiar with the procedures required to make the connection, productivity is enhanced, I
and there is less chance for error.
Some types of connections
ple, welding and post-tensioning.
require skilled craftsmen to accomplish,
The fewer of these skilled trades required, the more
for exam- I
economical the connection will be.
I
3.1.7.5 Standardize sizes of components. Whenever possible, such things as field bolts,
loose angles, etc., should be of common size for all connections.
for error, and the time required to search for the proper item.
This reduces the chance I
3.1.7.6 Use connections
pack, cast-in-place
that are not weathersensitive.
concrete,
Such materials as grout, dry-
and epoxies need special provisions to be placed in cold
I
weather. Welding is slower when the ambient temperature is low. If the connections are
designed so that these processes must be completed before erection can continue, costly
I
delays may result.

3.1.7.7 Use connections that are not susceptible to damare in handling.


I
Reinforcing bars, steel plates, dowels, and bolts that project from the precast piece will
often be damaged in handling, requiring repair to make them fit, especially if they are of
I
small diameter or thickness. It is often better to use items that are larger than required
by design, just so they will be rugged enough to withstand the rough handling often re- I
ceived. Anchor bolts for columns which project from cast-in-place foundations should be
at least 1 in. diameter so there is less chance of their being bent. Threads on projecting
bolts should be protected from damage and rust.
I
3.1.7.8 Allow alternates. As with fabrication, the precast concrete manufacturer or I
erector may prefer certain details or procedures not anticipated by the designer. Allow-
ing alternate
nections.
solutions will often result in more economical and better performing con- I
3.1.8 Appearance I
When the precast members are exposed, the appearance of the connection is
I
3.10
I
I
I often an important consideration. It is sometimes necessary to compromise fabrication

I and erection
ance.
simplicity, and hence, increase the cost, to provide a satisfactory appear-

I 3.2 Concepts of Connection Design

When a serious problem with a connection is encountered, the cause is often

I that the designer, fabricator,


or erector either failed to recognize, or failed to thorough-
ly follow through, a fundamental concept. Jn this section, some of these concepts are
explored.
I 3.2.1 Load transfer paths

I The purpose of a connection is to transfer load from one precast member to


another, or from a precast member to another element of the structure, or vice-versa.

I In most connections, the load will be transferred


tion by various mechanisms.
through several elements of the connec-

As an example, consider the steel haunch shown in Fig.34 To avoid penetrat-


I ing the form, the exterior portion of the haunch was welded on after the column was
removed from the form. The uniform load, W, must be transferred to the column by the
I various mechanisms below:
(1) Beam to bearing area by shear strength of the beam.

I (2) Bearing area to haunch through compression of the pad.

I
I
I >-T

I
I
I
Fig. 3-6 - md patha on connection
I
I 3.11
I
(3) Haunch to steel plate through shear and flexural strength of steel shape.
I
(4)
(5)
Through the plate via welds.
From embedded steel shape to column concrete through bearing. I
The tensile force (caused by volume change shortening) follows these paths.
(6) Concrete beam to reinforcing bars by bond. I
(7) Reinforcing bars to bearing angle through weld.
(8) Bearing angle to steel haunch through friction on top and bottom of the
bearing pad. Most of this volume change force is then relieved through
I
(9)
deformation of the pad.
There is still a minor amount of tensile force transmitted through the
I
(10)
welds to the steel plate, and then to the embedded shape.
This is then resisted by bearing on the projecting studs and by bond be- I
tween the embedded shape and the column concrete.

Each of these load transfers establishes the forces to be used in designing the
I
connection. It is obvious that the fewer the number of load transfers, the less chance for
error. Load transfer mechanisms and load transfer devices are discussed in Sections 3.3 I
and 3.4, respectively.
3.2.2 Failure modes I
In a manner similar to following load paths, the connection designer must exa-
mine each possible mode of failure in the connection and its component parts. In some of
I
the simple connections, the critical failure mode will be quite apparent. In others, it is
not as obvious, and the behavior often has only been determined by laboratory testing. I
An example of a component of a connection that has several possible failure
modes occurs when it is necessary to recess the connection
requires what is called a “dapped-end” beam and is illustrated
into the beam depth.
in Fig. 3-7.
This I
I
0
c
%- I
I
I
Fig. 3-7 - Possible failure modes in dnpped+nd beam
I
3.12
I
I
I The PC1 Design Handbook(3*2) lists five potential failure modes which should be
investigated in designing the dapped-end of a beam:
I (1) Flexure (cantilever bending) and axial tension in the extended end.
(2) Diagonal tension emanating from the reentrant corner.
I (3) Direct shear at the junction of the dap and the main body of the member.
(4) Diagonal tension in the extended end.

I (5) Bearing on the extended end.


A sixth potential failure mode is diagonal tension in the undapped portion (3.9).

I Design equations for investigating


When investigating
each of these failure modes are given in Sect. 3.5.3.
the various possible failure modes in a connection, the type
failure should also be determined, i.e., ductile or brittle. As a general rule,
I of potential
a higher load factor should be applied to those potential failure modes that are the most
brittle.
I 3.2.3 Stability and equilibrium

I Some of the most common problems in precast, prestressed concrete structures


are caused by the failure of the designer or constructor to consider stability and equih-
brium of the structure and its components, not only in its completed state, but during all
I phases of construction.
A typical example is the case of a ledger or L-shaped beam as shown in Fig. 3-
I 8. Because of the eccentric loading, the beam is subject to torsion and tends to “roll”, or
rotate on its supports. It is not unusual for the designer to go to great lengths to assure

I that such beams are adequately


the end connection requirements
reinforced to resist torsional stresses, and then neglect
to restrain the beam against torsional rotation.

I
I
I
I Support

I
I
I 3.13
I
In some structures, restraint against torsional rotation in the completed struc- I
ture is only provided by in-situ concrete which is placed after the precast members are in
place. Jn these cases, temporary restraint must be provided during erection. This is I
frequently very difficult and costly to do. It is better if connections are designed which
will make the members torsionally
the final condition.
stable during all phases of construction, as well as in I
In most precast, prestressed concrete structures, it is desirable that permanent
lateral stability be provided by shear walls or cross-bracing, rather than by moment
I
resistant frames. Moment-resisting connections are complex, congested and expensive.
Also, volume change restraint forces build up severely in continuous multi-bay and multi- I
story frames.
Lateral forces are distributed to the stabilizing assemblies through diaphragm I
action of the floor and roof decks. Since the structural frame must be placed before the
decks, temporary stability must be provided and constantly
buildings this may require a detailed
maintained. In multi-story
analysis and careful planning for all construction
I
phases. Sometimes connections are designed which will provide some temporary moment
resistance during erection, and then “released” (by removing bolts, or cutting welds loose)
I
when the permanent
encouraged, because
lateral stability
it is too easy for
assemblies sre in place.
the constructor
This practice
to forget to release
is not
the I
mechanism, and unplanned stresses could result.

3.2.4 Stress Relief Measures I


ln many cases it is far more desirable to design connections
total restraint of a precast prestressed member, than to attempt to resist the stresses
that will prevent
I
caused by this restraint. Examples of loadings for which relief is desirable are severe
earthquakes and volume change shortening. I
Connections which provide restraint relief cannot also provide resistance to
lateral loads. Jn general, it is desirable to rigidly connect no more shear walls or frames
than absolutely necessary to resist the lateral loads. AR other connections should be
I
designed for relief of restraint.
I
3.2.4.1 Elastic and inelastic strain. For severe earthquake, the most common relief
measure is through inelastic strains-yielding of steel and cracking of concrete, i.e., duc-
tility. This is discussed in more detail in Sect. 3.1.2 and Part 2.
I
Elastic and inelastic strains are also useful in relieving
For example, in Fig. 3-9 the angle at the top of the beam would be effective
other types of stress.
in resisting
I
I
3.14
I
I
I
I
I
I
I
I pig. 3-9 - Yielding top connection Fig. 3-10 - Slotted connection

I torsional rotation but would yield enough under gravity loads so that substantial negative
moments would not be attracted, even if compression is developed at the bottom.

I 3.2.4.2 Bearing pads. To relieve the restraint


and temperature
of shortening caused by creep, shrinkage
change, flexible pads are normally used under the bearings of beams and

I stemmed deck members.


ness.
These pads relieve restraint by deforming within their thick-

I There are several suitable materials and combinations


used as flexible bearing pads. These are described in Sect. 3.4.8.
of materials that are

I 3.2.4.3
allowed
s Stresses can also be relieved if certain portions of the connections
to slip under movements caused by volume changes, or in some cases, earth-
are

I quakes.
Some types of bearing devices are designed to provide slip characteristics. The
most common are thin plastic or hardboard bearing strips used under hollow-core slabs.
I Tetrafluorethylene (TFE - trade name Teflon) bearing devices are also used where very
low friction is desired.
I It is common practice to use slotted holes in clip angles as shown in Fig. 3-10.
When properly placed, this will allow some horizontal slip, but will restrain torsional

I rotation. However, the slot is also used for field adjustment, and all too often the bolt
ends up in the end of the slot, preventing further movement. There is also a tendency for

I workmen to tighten the bolt too much. If the purpose of the slot is to relieve restraint,
erection instructions should be clearly indicated and carefully inspected.

I
I 3.15
I
3.2.5 Expansion joints. The term “expansion joint” is applied to joints which extend
I
completely
dering lateral
through the building, effectively separating it into two structures
The PC1 Design Handbook (3.2) , Sect. 4.3, contains a discus-
movements.
when consi-
I
sion of expansion joints in precast, prestressed structures.
A true “expansion joint” is only required if the movements resulting from tem- I
perature rise are greater than the shortening caused by creep and shrinkage. ln precast,
prestressed concrete
contraction are needed
buildings, this rarely, if ever, happens.
to relieve the strains
Instead, joints that permit
caused by the additive effects of
I
temperature drop, creep, and shrinkage.
There are conflicting opinions regarding the spacing of expansion joints. These
I
joints are frequent sources of problems and often require maintenance, so it is desirable
to have as few expansion joints as possible. I
The National Academy of Science’s publication, “Expansion Joints in
Buildings”(3*10) provides guidelines for determining expansion joint spacing and designing I
the joints. ‘lhe recommendations in that report are based on a study of government
buildings and analytical study of the effects of uniform temperature change on typical I
two-dimensional elastic frames. This study found the need for expansion joints to be a
function of the following parameters (in addition to the length of the building).
1) Framing material (concrete or steel)
I
2)
3)
Configuration of the building-rectangular, L-shaped, T-shaped, etc.
Whether or not the building is heated and/or air conditioned
I
4) The base fixity condition of the columns
5) The relative stiffness against lateral displacement along the length of the I
building.
The above report did not address precast concrete buildings with “soft” connec- I
tions, i.e., those which use bearing pads or other restraint relieving measures discussed
previously. Experience of others has shown that if all or most connections are “soft”, the
distance between joints can be substantially increased over those recommended in Refer-
I
ence 3.10 and 3.2.

3.3 Designing Load Transfer Mechanisms


I
Since the function of connections and their components is to transfer loads, it is
necessary to understand the various load transfer mechanisms, and how to analyze
I
them. How these mechanisms interact within a connection was discussed in Sect. 3.2.1.
In this and the next section, these transfer mechanisms will be examined in detail.
I
Where applicable, design methods will be given.
I
3.16
I
I
I For the design equations in these sections, the forms of the AC1 Building
Code(3.11) will be used as much as possible. Jn 19’7’7,AC1 introduced a new system of
I notation to be used in strength design of reinforced concrete. Previously, the
subscript “u” denoted either the applied factored forces (MU, Vu, Pu, etc.) or the design
I strength (termed “ultimate strength” prior to 1971). In the 1977 edition of the
Building Code, the subscript ‘%I” denotes only the applied factored forces. The subscript
I “n” denotes the “nominal strength”. The “design strength” (or “ultimate strength”) is the
nominal strength multiplied by the strength reduction factor, + , for example
I +sl, evn9 0% * Thus the design of a member or component requires that

$ kpg, vu =s+v,,Pud 4% , etc.


I
In presenting equations for strength design, this leads to the apparent algebraic
I redundancy of having the term $, on both sides of the equation. However, to avoid the
inadvertent neglect of the $ -factor, the equations presented here are in terms of “de-
I sign strength” rather than “nominal strength”.
The AC1 Building Code has also presented all of the design equations in force
I rather than unit stress terms. This sometimes leads to unnecessarily complex relation-
ships, and can be somewhat confusing, especially when comparing one design equation

I with another. Thus, this report will use unit stresses in the design equations when it is
considered appropriate.

I 3.3.1 Bearing
Nearly all connections of precast, prestressed concrete involve the bearing

I strength of concrete. The AC1 Building Code(3’11) limits the unit design bearing stress
to:
fbu = 0.85 +f;$& (3-l)
I where:
fbu = limiting design bearing stress, psi
I
4 = 0.70
I f; = specified 29 day compressive strength of
concrete, psi (based on cylinder tests)

I Al = loaded area, sq in.

I
I 3.17
I
= maximum area of the portion of the supporting
I
surface that is geometrically similar to and
concentric with the loaded area, sq in. I
Rese&,(3.1fL, 3.13)
I
has established that bearing failures of plain concrete are
usually characterized by a splitting of the concrete perpendicular to the applied load.
Thus, the limit on bearing stress was determined to be a function of the splitting tensile
I
strength, rather than the compressive strength of concrete. The splitting tensile
strength has, in turn, been established to be a function of the square root of the compres-
I
sive strength.
Further, the AC1 equation does not consider the effects of a tensile force I
perpendicular to the applied load. This was found to substantially reduce the bearing
strength of the member. I
Kriz and Rathsc3’12), in a series of tests on column heads, mostly with lateral
reinforcement, suggested the following equation: I
Nu’vu
fbu = (69a y;, x (1 + C1,&%(Cr/80) (3-2) I
where (See Fig. 3-11):
S = distance from center of load to free edge of member, in.
I
w = width of bearing, in.
Cl = Owhens (2 in.; 2.5 when s? 2 in. I
Avf = area of reinforcement perpendicular to splitting crack, in.
b = length of bearing, in. I
c, = the product of SW,but not more than 9.0, sq in.
I
I
I
I
Fig. 3-U- Bearing on plain concrete
I
3.18
I
I
I N, = factored tensile force parallel to plane of bearing, lb.

I V” = applied load perpendicular to plane of bearing, lb.


The Kriz and Raths tests were somewhat limited in scope, particularly regard-
ing the effects of tensile force (Nu) on plain concrete specimens. Therefore, on reex-
I amination of the data, Raths proposed the following equation, with the concurrance of
the Prestressed Concrete Institute Committee on Connection DetaiLs, for inclusion in
I Reference 3.3 for bearing on plain concrete:

I
fbu =
I This equation was also published in the PC1 Design Handbook(3Sz)and has been
used successfully by the precast, prestressed concrete industry in the U. S. since about
I 1971.
References 3.2 and 3.3 also suggests that when the limits prescribed above are
I exceeded, reinforcement can be designed by shear-friction (see Sect. 3.3.2.1).
In a much more comprehensive study, Williams (3.13) compared the results of
I some 15 research studies, including those of Kriz and Raths, and also conducted some
very extensive testing. A plot of all the data from Williams’ tests and the other studies

I referenced suggested a best fit curve of:

% -0.47 (3-4)
fbu = 6.92 fct (-)
I %
where:
I f,t = splitting tensile strength of concrete, psi.
ACl(3.11) suggests that when values of fct are not determined from test, the

I quantity 6.7X&A may be substituted for f,,, where i is a coefficient used with light-
weight concrete. Thus a suggested design equation based on Williams data would be:

I fbu
= $45x -f22
l- A1 (3-5)

I Based on the results of 12 tests, which included a horizontal (parallel to bearing


plane) component, Williams suggests the following relationship:
I fbuh fbu Nu (S-6)
-=,+T
f CU f
I
I 3.19
I
where:
I
fbuh
f C”
=
=
bearing strength with horizontal
characteristic cube strength
component
of concrete, based on 152 mm
I
fbu =
cubes
bearing strength without horizontal component I
m = a coefficient which varies from -1.52 to -2.27
Combining Eqs. 3-5 and 3-6, and accounting for the difference between cube I
strength and cylinder strength, a suitable design equation which would reasonably fit
Williams’ data would be: I
(3-7)
I
@=
where:
0.70 I
X= 1.0 for normal weight concrete, 0.85 for “sand-lightweight” concrete, and 0.75
for “all-lightweight” concrete as defined by AC1 318-77. Other terms as I
defined for Eqs. 3-1 and 3-2.

tion of v%&
The data from Reference 3.13 indicates that if Eq. 3-7 is used, the AC1 limita-
s 2 does not apply. However, because of the limits of values tested, Fq.
I
3-l should be used as an upper limit.
fbu hex) = (0.85 $f;) 2 = 1.2 f;
Thus:
I
(3-S)

3.3.2 Shear Strength


I
Many connections achieve load transfer through the shear strength of either the
concrete or steel components of the connection.
I
3.3.2.1 Concrete shear strength
I
The shear strength of concrete has been, and continues to be, a much research-
ed, yet still controversial subject. Most of the current provisions for shear in AC1 318-
77(3.11) are the result of the work of Joint Committee 326 (now 426) of the American
I
Concrete Institute and the American
in three parts in 1962(3*14).
Society of Civil Engineers.
Since that time, additional
reports of the Joint Committee
Their report appeared
I
have appeared13’15), and AC1 pilblished a set of papers in 1974(3’16). This later work,
while often critical of the AC1 Code provisions in the 1963 and 1971 versions of AC1 318, I
has resulted in only minor changes.
I
3.20
I
I
I It is universally agreed that so-called “shear failures” in concrete are actually

I caused by diagonal tension failure. Thus, while classic reinforced concrete design as-
sumes that concrete has no tensile strength, the “shear strength” of concrete is, in fact,
dependent on the tensile strength.
I Most of the formulas that limit shear stress in AC1 318-77(3*11) and other
building codes are derived from the principal stress equation:
I
ft (max) = l/2 ft +Aft/2)2 + v2 (3-9)
I where:

I ft (max) = maximum principal tensile stress


ft = tensile stress caused by flexure and axial tension
= direct shear stress on the member
I V

The limiting value of ft (max) of 4 q for normal weight concrete appears to


be substantiated by tests.(3’17). The only place in AC1 318-77 where principal tensile
I stress is directly mentioned (Sect. 11.4.2.2) specifies this value (with appropriate
strength reduction factors and modifications for lightweight aggregate concretes) as the
I limit.
The various building codes and research distinguish between ‘beam shear” and

I “punching shear” in concrete members. In connection design, ‘beam shear” is the mech-
anism that transfers loads from flexural members to connections. For full-depth mem-
bers, the shear design procedure is well defined by AC1 318-77 and other building codes.
I For special end conditions, such as the dapped end described in Sect. 3.2.2, special de-
signs are necessary.
I In the codes, fpunching shear” is used in the design of slabs and footings to
describe the shear failure mode of a concentrated load or reaction that bears on the
I lab. AC1 318 limits the l’punching shear” stress on an unreinforced section to 4fi;
The surface area to which this limiting stress is applied has been somewhat arbitrarily
.

I chosen as the perimeter of an area at a distance of d/2 from the load area, times the
effective depth of the slab. (See Fig. 3-12). The British Code(3.18) prescribes a larger
perimeter, but uses a correspondingly lower shear stress. ‘Ihis has been defended, be-
I cause tests have shown that the “punched-out” surface is nearly always considerably
larger than the 45’ failure plane that is the basis for the AC1 definition.
I A phenomenon similar to punching shear occurs in precast concrete connections
when inserts, studs, etc. are subjected to a tension force. Design procedures for these
I items are discussed in Sect. 3.3.4.

I 3.21
I
I
I
I
I
I
I
Fig. 2-12 - Comparison of ACI and British criteria
for pullching shear I
In the 1962 Joint Committee report (3.14) the following statement is made: I
II. .. the goals of a complete understanding and of a fully rational solution to
the problem of computing diagonal tension strength have not been attain- I
ed. In view of these circumstances, it appears necessary at this time to
abandon the classical procedures in favor of a logical, though empirical
approach which takes into account the major variables affecting diagonal
I
tension strength as shown by test results.”
In 1980, the “goal of complete understanding” does not seem to be any closer.
I
Thus, the design of connections which depend on the strength of concrete in shear and
tension to transfer loads should be verified by test data. Unfortunately, well-document-
I
ed test data in this area are scarce.
3.3.2.2 Shear strength of steel
I
The shear yield stress of structural steel has been variously estimated as be-
tween l/2 and 5/8 of the tensile yield stress(3.20). However, the AISC
I
Specification(3*1s) allows a value equal to 2/3 the recommended basic allowable tensile
stress, substantially the same as it has been since the first edition published in 1923.
I
‘This apparent reduction in factor of safety is justified by the minor consequences of
shear yielding and by the effect of strain hardening, as well as the experience of over 50 I
years.
I
3.22
I
I
I Shear strengths of fastening devices, such as bolts and welds, are given in

I Reference 3.19. The design of these fasteners, as they apply to connections of precast,
prestressed concrete structures is discussed in Sect. 3.4.

I 3.3.2.3 Shear-friction
The shear friction hypothesis was introduced by Mast in 1968(321). The con-

I cept provides a rational approach for the design of reinforcement across a plane that is
subjected to shearing forces. This plane can be a crack, (real or potential), or an inter-
face between two materials, such as steel and concrete, or concrete cast on previously
I hardened concrete.
Simply stated, the shear friction concept states that slip along the plane is
I prevented by “friction” in a manner analogous to the classic friction formula:

I F =pN (3-10)
In this case P is the design shear strength, Vn, that can be resisted,u is the shear-fric-

I tion coefficient and N, the “normal force”, is the strength of the reinforcement, Avf,
across the crack. In order for the two surfaces of the interface to part and slip, the rein-
forcement crossing the plane must yield. Substituting AC1 notation into Rq. 3-10, yields:
I
Vn=u A f 3 vU
- (3-11)
I VfY $

or for design purposes:


I vU
ptf =‘fy
I The shear-friction idea was quickly adopted by the design community for a
variety of connection design situations, and appeared in the 1971 AC1 Building Code, the
I first edition of the PC1 Design Handbook published in 1971 and in the PC1 Connections
Manual(“11.
I In those publications, the value ofu was set at 1.4 for an assumed crack plane
in concrete cast monolithically, 1.0 for concrete cast against hardened concrete and 0.7
I for concrete placed against steel.
In the subsequent proliferation of the use of the shear-friction concept, it was

I found that for many cases, the amount of reinforcement required to satisfy the equation
was very high, and appeared to add unnecessary costs to precast, prestressed members.
Both Mattock(3’22s 3.44) and Raths(3’23) observed that the shear resistance was a func-
I
I 3.23
I
tion of not only the force in the reinforcement, Av f and the interface condition,u ,
I
fY
but also the area of the interface. For monolithic concrete, Mattock suggested an
approach he called “modified shear friction”, in which:
I
V, = 0.8Avffy + AcrK
where:
o-12)
I
A”f = Area of reinforcement nominally perpendicular to the assumed crack
plane, in2 I
fy = Yield strength of Avf, psi (equal to or less than 60,000 psi)
A cr = area of the crack interface (varies depending on the type of connection) sq
in.
I
K = 400 psi for normal weight concrete, 200 psi for “all lightweight” concrete,
and 250 psi for “sand-lightweight” concrete.
I
In Rq. 3-12, the fit term represents the contribution of friction to shear- I
transfer resistance, and the second represents the contribution of the resistance to
shearing off of protrusions on the crack faces and dowel action of the reinforcement. I
The Raths(3’23) approach was to vary the shear-friction coefficient with the
area of the interface. The Prestressed Concrete Institute Committee on Connection
Details assigned the task of investigating the different approaches to committee member
I
A. P. Shaikh of the University of Wisconsin-Milwaukee.
In his study, Shaikh also included some earlier work by Birkeland (3.24). He I
found the three different approaches to produce almost the same results(3’25) and rec-
ommended, with the concurrence of the PC1 Connection Committee, the Raths “effec- I
tive shear-friction coefficient”, because of its similarity to the by-now established shear
friction method. I
The second edition of the PC1 Design Randbook(3’2) used the Connections
Committee’s recommendations, changing only the form to make them consistent with
AC1 318-77 in terminology. The equations as shown in Ref. 3.2 are reprinted here:
I
where %f = $i+
I
e

$ = 0.85 I
v = Applied factored shear force, parallel to the assumed crack plane, lb
up
1000 i2Acru I
‘e= V (3-13a)
up
I
3.24
I
I
I lbble 3-l - Shear-friction coefficients
Crack interface Recommended Maximum Vup, lb
I condition !J
1 Concrete to concrete,
I cast monolithically
2 Concrete to hardened concrete
1.4 0.30X2 f; AcrS1000X2 A0

I with roughened surface


3 Concrete to concrete
1.0 0.25X2 f; A0 *1000X2 A0

smooth interface 0.4 0.15 h2 f; A0 4600 A2 A0


I
4 Concrete to steel 0.6 0.20X2 f; Acr d 800X2 Acr
I
I x
=
= 1.0 for normal weight concrete
(f,.,/6.7) qfor sand-lightweight or aII lightweight concrete, but not
I greater than 1.0. If fct is unknown:
0.85 for sand-lightweight concrete
0.75 for aII-Iightweight concrete
I f ct = splitting tensile strength of concrete, psi
u = values from Table 3-l
I (other notation same as Eq. 3-12)
The shear-friction reinforcement should not be less than:
I (3-14)

I unless A,,f calculated by Eq. 3-13 is increased by one-third.


When axial tension is present, additional reinforcement area should be provided:
I
(3-15)
I
The constant “1000” in Eq. 3-13 is somewhat controversial(3.26). The University
I of Washington tests were performed on precracked specimens. Using a “mean value” of
those tests, Mattock suggested that 900 was more representative (3.23) The higher value
was based on a lower bound curve using untracked specimen data (3.25(
I AC1 318-77t3*11) only recognizes the method shown in Eq. 3-11. At this writing
proposals for change by 1983 are being considered.
I
3.25
I
I
Mattock states that in using Eq. 3-12, the value A,,ffy/A,r should not be less
I
than 200 psi. ln Ref. 3.25, Shaikh suggested this value could be lowered to 120 psi,
resulting in Rq. 3-14. However, this limitation was rendered essentially ineffective when I
the statement “unless Avf is increased by one third” was added by the PC1 Connections
Committee. I
The shear-friction concept has been applied to the in-plane shear transfer
between
reported
hollow-core
on full-scale
slabs used as lateral
tests using the concept.
load resisting diaphragms. Moustafa(3’56)
‘lhe results appear to indicate that the
I
limitation of F.q. 3-14 is too severe, but increasing
3-13 by only one-third is unconservative.
the reinforcement provided by Eq.
Further research in this area seems warranted.
I
3.3.3 Tensile Strength
I
The tensile strength of steel and concrete is frequently used as a load transfer
mechanism. The strength of concrete
crete its shear strength,
in direct tension is the property
as discussed in Sect. 3.3.2.1.
that gives con-
It need only be added here that
I
tensile (shear) failures of concrete
with substantial factors of safety.
are brittle and sudden and should, therefore, be used
I
On the other hand, since most steels are very ductile, it is desirable to design
connections so that tensile yielding of steel occurs before concrete crushes or fractures. I
3.3.4 Anchorage

When a tensile load is transferred between concrete and a steel device, the
I
steel must be anchored to the concrete.
device-reinforcing bar, stud, bolt, etc.
The method of anchorage depends on the type of
I
Anchorage failures are usually sudden and brittle in nature. It is therefore
usually recommended that anchorage strength be sufficient to force failure of the steel I
connecting device.

concrete.
Reinforcing bars and prestressing strand are usually anchored by bonding to the
Required lengths for anchorage are specified in AC1 318(3’11). Very often,
I
there is insufficient bond length available to anchor the bars, and supplemental mechan-
ical anchorage is required. This can be accomplished by bar hooks or welded cross-bars
I
as shown in Fig. 3-13. Most other types of steel devices depend on mechanical anchor-
age.
I
3.3.4.1 Anchorage of bolts, studs and inserts

Studs and bolts depend on the head to engage a concrete area that causes a
I
cone-type pull-out failure as shown in Fig. 3-14. Similarly, inserts which receive bolts or
I
3.26
I
I
I
I
I
I
I
I Fig. 3-14 - Assumed failure surface for
Fig. 3-13 - Mechtmlcel anchorage of reinforcement headed studs and bolts
I threaded rods employ various types of anchorage devices to try to force this cone type of
failure (Pig. 3-15). The PC1 Design Handbookf3”) and other publications (3.27, 3.28, 3.29)
I show design methods for this type of cone failure that assumes a shearing surface and
applies a unit shearing strength to that surface area.
I In these design methods, it is assumed that the pull-out cone makes an angle of
45’ with the surface of the member, and the unit shearing strength applied Is 4$ Jf;!,

I the same used for punching shear in slabs. This value has apparently been verified by
tests which are cited in references 3.27 through 3.29. In most direct pull-out tests, a
somewhat flatter shear cone has usually been observed, although the ultimate strength
I agrees closely with strength calculated using a 45’ cone and a unit stress 4 q . As in
punching shear, this suggests that a more precise design could be based on the flatter
I shear cone (more surface area) and a lesser unit stress. Nevertheless the presently used
criteria has had good experience for over 10 years.
I
I
I
I
F&3-15-shear-cone
I
I 3.27
I
When embedded steel devices are placed near the edge of a member, or when
I
two or more are used together, and spaced so that the full shear cone cannot develop,
special procedures are necessary for design. In this case, reference 3.2 recommends that I
the shearing surface area be assumed to be a truncated pyramid as shown in Fig. 3-16.
Reference 3.8 cautions that embedded devices in thin members can develop an even I
smaller shearing surface which extends through the member as shown in Fig. 3-1’7.
While these references are illustrated for steel studs, they apply also to other
types of devices that depend on mechanical anchorage.
I
An area that has apparently not been explored is the effect of steel reinforce-
ment passing through the assumed shear surface as shown in Fig. 3-18. It appears that
I
I
I
I
I
I
I
I
I
I
I
I
l’@.3-16-Pull-autsurfaceainstudgroups
I
3.28
I
I
Rd”tCWC-“l

II I
= = = = ,.:::,:: i zzz==
;; <’ ,:
--- Jo...::.
::.. ,i. = zrz==
0
II II F.YIIYle sm.ce
G

TENSION MOMENT

F@. 3-17 - Stud groups in thin sectiorrs Fig. 3-18 - Reinforcement placed

I subject to tension and moment tkough failure cone

the principles of shear-friction (Sect. 3.3.2.3) apply, but this has not been verified by
I test. Such reinforcement would almost certainly improve ductility.
Studs, inserts and embedded bolts are frequently subjected to shearing forces

I (Fig. 3-19). Reference 3.2 indicates shear strength to be controlled


the steel connector, unless the embedded item is loaded toward a free edge. However,
by the strength of

some tests on studs have indicated that concrete strength may be more critical (3.30).
I This is particularly so in concrete with strengths below 5000 psi, and when lightweight
aggregate is used. Reference 3.29, based on the data from Reference 3.30 suggests the
I shear strength be limited to:

I $A’, = 6.66 $Abf;0’3 Eco*44

I
I
I
\ h
I
E&3-lS-Shcxwloadingonastud
I
I 3.29
I
where I
v, =shear strength of the concrete in lbs.
4
Ab =
=0.85
area of the stud, sq in.
I
f;
E, =
= compressive strength of the concrete, psi
modulus of elasticity of the concrete, psi
I
A review of the data from Ref. 3.30 suggests that the following equation could
also be used: I
@Vc= 900 l$A q (3-17) I
where
x =
1.0 for normal weight concrete
I
=
0.85 for sand-lightweight concrete
=
0.75 for all-lightweight concrete
I
When the shear load is in the direction of a free edge, the shear strength of the
concrete becomes more critical. References 3.2 and 3.28 specify the following limita- I
tion:
$V, = 3250 C$(de - 1) J- -f;5000 (3-18) I
Ref. 3.29 recommends that the shear strength calculated from equation (3-16)
be multiplied by the factor:
I
d -1
e (3-19)
I
8dS

In Eqs. (3-18) and (3-19): I


de = distance from free edge
d, = diameter of the stud I
Equations (3-18) and the combined (3-16) and (3-19) represent the viewpoints of
two different authors and are based on the same test data (3.31). I
There has been very little similar research with connectors other than studs.
Limited tests on inserts in low-strength concrete ( f& = 3000 psi) led to the following
equation printed in Reference 3.2:
I
s$Vc= +(2500de - 3500) (3-20)
I
I
3.30
I
I
I This equation is probably conservative for the higher strength concretes typical

I of precast, prestresscd concrete construction. The absence of data, however, precludes


the use of other equations. Reference 3.2 states that if the insert is located 4 times the
insert embedment from the edge, the shear strength can be assumed equal to the pull-out
I strength.
In most applications of embedded devices in precast concrete connections, a

I combination
be used:
of shear and tensile force is applied. For desk&an interaction equation can

I (3-21)

I where:
= applied factored tensile and shear forces, respectively
I P”, Vu
@PC, +vc = design tensile and shear strengths, respectively
(based on concrete strength)
I A minor controversy exists over the value of the exponent, K. References 3.3,

I 3.29, and 3.31 use K = 5/3. ln references 3.1, 3.2 and 3.28, K = 4/3. Both are based on
the research at Lehigh University (3.31). Examination of the data indicates that 5/3

I exponent represents more or less average values of test results, while the 4/3 exponent
produces a lower bound curve(3*8).
Note that all of the design equations in this section are based on the limitations
I of the concrete. Design strengths of the various steel devices are given in Section 3.4.
For combined tensile and shear strength, all of the above cited references agree on the
I interaction equation:

I (3-22)

where
I 4%) eVs are the design strengths based on steel capacity

I 3.3.5 Friction

In connections of precast, prestressed concrete elements, lateral loads are


I frequently transferred by friction. Jn some cases this may be desirable and planned, as
for example, in the transfer of forces caused by wind loads from a floor or roof dia-
I
I 3.31
I
Table 3-2 - Static coefficients of friction of dry materiad3’2)
I
Material I
Elastomeric to steel or concrete 0.7
Laminated cotton duck fabric
to concrete 0.6 I
Concrete to concrete 0.8
Concrete to steel
Steel to steel (not rusted)
TFE to TFE
0.4
0.25
0.05”
I
Hardboard to concrete 0.5
Multipolymer
to concrete
plastic (non-skid)
1.2Q'
I
Multipolymer plastic (smooth)
to concrete
I
I
phragm to a shear wall in a bearing wall structure. In other cases, friction can cause
restraint of volume change forces which may be undesirable. I
Table 3-2, taken from the PC1 Design Handbook t3m2)gives static coefficients of
friction for various interface conditions
purposes, these values should be assumed maximum,
frequently found in connections. For design
and thus should only be used for
I
assessing the undesirable effects of friction.
Reference 3.2 recommends that if friction is depended upon to resist loads, the
I
coefficients of friction in Table 3-2 be divided by five. Many designers consider friction
too unreliable to depend on for load resistance, and require at least nominal reinforce-
I
ment or other mechanical ties in all connections (3.7).
3.4 Load Transfer Devices
I
performance
A variety of devices are used to transfer loads in connections.
of connections reduces to the design and performance
‘lhe design and
of each load transfer
I
device employed.

3.4.1 Bolts and Threaded Connectors


I
Various types of bolts and other threaded connectors are used in connections of
precast, prestressed concrete. The biggest advantage of these devices is that they can
I
be quickly assembled and erected. ‘lhe primary disadvantage is that close tolerances are
required for the placement of the connector and its receptacle when embedded in a
I
concrete member.
I
3.32
I
I
I In most connections, the bolts are shipped loose to the site of final erection, and

I are threaded into receptacles cast into the concrete (See Sect. 3.4.2, “Inserts”). Occa-
sionally a precast concrete member will be cast with a threaded connector projecting
from the face. This is usually an undesirable practice because such items are vulnerable
I to damage during handling. Also, unless the projection is from the top of the member as
cast, stripping from forms is extremely difficult-impossible with some types of forms.
I The types of threaded connectors that are most commonly used include: 1)
standard bolts; 2) high-strength bolts; 3) threaded steel rods; 4) coil bolts and coil rods.

I Other proprietary connectors are also available.


3.4.1.1 Standard bolts

I Standard bolts as defined here are those meeting ASTM A307. Threads are
“Coarse Thread Series” as specified in ANSI B1.l, and shown in Table 3-3.
I Design of standard bolts is in accordance with the AISC Specifications (3.19),
which requires working stress design using unfactored loads. The allowable stresses are

I shown in Table 3-4.


When bolts are used in connection design with factored loads, a reasonable
approximation can be made by multiplying the allowable working stress by 1.65. This can
I be used assuming a’ capacity reduction factor, o , of 1.0.
3.4.1.2 High-strenth bolts
I High strength bolts (ASTM A325 or A490) were developed for friction-type

I connections between structural steel members. They have more than twice the tensile
stress capacity of A307 bolts. Their application requires controlled tensioning of the
fastener to develop sufficient force to prevent slipping of the connected parts. To take
I full advantage of high strength bolts in friction-type connections they must be tightened
using calibrated torque wrenches or load indicating washers.
I Because of creep and minor crushing, it is questionable as to whether a high
strength bolt will hold its tension when tightened against concrete. Thus, high strength
I bolts are used infrequently in precast concrete connections. Some applications and
limitations are shown in Fig. 3-20.

I 3.4.1.3 Threaded steel rods


Steel rods of standard sizes and grades are sometimes used in precast concrete
I connections. The most common application is for anchor bolts at column bases. Allow-
able working stress for rods of ASTM A-36 steel are given in Table 3-4.

I
I 3.33
Table 3-3 - Screw thread, bolt head and nut standards
I
I
I
I
01 Bolt Width Width Height Width Width Height Width Widfh high,
F
I”.
c
I”.
H
I”.
I
~
716 1 5116
1.1116 1.114 7116 I
1-l/4 l-7116 112

1 I-112 2-116 1,116 1.112 1-11116 1,116


II-7116 I-11116 9116
l-516 I-71 518
I
I-114 I.716 2-516 716 I-718 2.116 716 2 24116 716

I
I
I
N”, Wi;th Wt$h He;ght WY Wph Height Wl;th Wph t+e$ht Wiph WFh Heip
size
I”. I”. I”. I”. I”. I”.
N
I”. tn. I”. I”. I”. tn. I”. I
I
I
Diameter 1 Ama
I
Th’dn
1%

MINtMUM LENGTH OF THREAD ON BOLTS
ANSI 816.2.1 - 1966
I
I
Diameter of Bolt. D, Inches
112 .4w ,196 ,126 ,142 13 Length 01 Bolt
112 516 34 716 I lW4
11
3.4 ,626 ,442 ,302 334 10 To 6 in. Inc. l-l,4 I-112 I.314 2 2-l/4 2414
718
1
516
I.114
.731
,836
507
.Wl
,765
307
,419
,551
,202
,462
.w
,226 1
9
6
over 6 I”. 1.112 ICY4 2 2.114 2.112 3
I
I.064 1.227 ,690 ,969 7

I
3.34
I
I
‘hble 3-4 - Allowable workiw stresses on threaded fasteners

I NOM INOM. TEN. ALLOWABLE STRESSES


NAL NAL SILE

I
I
BOLT AREA STRES: Fy = 36 A 307 A 325F A 325N A 325X
SIZE AREA I I I I I I I I
F, ( F, 1 F, 1 F, 1 Fv 1 F, 1 Fv I F, 1 Fv
I 4
in. I”.’ ill.’
THREAD GROSS
PART AREA
BEAR.
ING
TYPE
GROSS
AREA

THREAC ,

I MA
in.
ALLOWABLE FORCES = kios
-
I l/2”
(0.40)
(I.196 0.142 3.12 4.31 2.1 2.64

SW I1.3066 0.2260 4.97 6.75 3.3 4.52

I (0.507)
314” I I.4416 0.3345 7.36 9.72 4.6 6.69 13.25
(0.620)

I 719” ( MO13 0.4617 10.16 13.23 6.5 9.23

I1
(0.731)
1” ( I.7654 0.6057 13.33 17.26 a.5 12.11 23.56

I (0.636)
1.114”
(1.064)
11.2272 0.9691 21.32 30.00 13.3 19.36 36.62
- - i

I F = Friction
N = Bearing
type connection
type connection with threads included in shear plane

I X = Bearing type connection with threads excluded from shear plane

I
I
I
I
I
I
I
I 3.35
I
3.4.1.4 Coil bolts and rods
I
The threads on coil bolts and continuously threaded coil rods are extremely
coarse (Fig. 3-21). They are designed to fit the contour and diameter of the helically
I
wound coil of wire used in some inserts (See Sect. 3.4.2).
Coil bolts were originally designed for use in temporary connections, usually in
I
lifting devices for site cast (tilt-up) or plant cast concrete members. They are ideal for
use in and around concrete, because the thread is not easily clogged or damaged. Coil I
bolts are frequently used in permanent connections in precast, prestressed concrete
construction. The coarse thread allows fast action, and does not require extremely close I
tolerance in placement of the insert. This helps to speed up erection.
Continuously threaded coil rods (Fig. 3-21) are available in lengths up to 10 ft.
They have many useful applications in precast concrete connections. For example, the
I
connection between a cast-in-place slab or topping and a precast beam or wall panel.
One end can be threaded into a coil insert in the precast piece. The threads serve the
I
same function as deformations on reinforcing bars to bond into the cast-in-place con-
crete. This bond development length of coil rods is assumed the same as deformed I
reinforcing bars but this has not been verified by test.
Coil bolts and rods are available from several concrete accessory suppliers, but I
are not covered by standard specifications. Manufacturer’s catalogs give
recommended ultimate and working strengths. References 3.1 and 3.2 have tabulated
typical values, which are shown in Table 3-5. These values are based on static loads.
I
The performance of coarse threaded coil bolts under cyclic loading is suspect.
Experience with bolts used in lifting inserts shows that they wear after several reuses.
I
This may indicate that the bolts could loosen after several repeated load cycles.
I
Table 3-5 - Capacity of Coil Bolts and Threaded Coil Rods
I
Bdt Diameter Min. Coil
Penetration lid
Tensile
Strength IPJ
I Shear
Strength W,)
I I
1
l/2
314
1 l/2
2
2 l/2
13,500
18,470
37,870
8,100
11.080
22,720
I
1 II4 2 l/2 54.960 32.980
1112 3 83.340 50,000
I
I
3.36
I
I
I
I
I
I
I COIL BOLT

I F&.3-5x-coilboltsandcoilrods

I 3.4.2 Inserts

I A wide variety of inserts are commercially available. Typical examples are


shown in Figs. 3-22 through 3-25. Basically the inserts vary only in the receptacle used
to receive the connector and in the manner in which they are anchored to the concrete.
I This discussion does not include the wedge inserts shown in Fig. 3-25. Wedge
inserts are sometimes used to connect cladding panels to steel frames, but are unreliable
I for most structural uses.
3.4.2.1 Receptacles in inserts
I Inserts which are cast into the concrete use one of three basic types of recep-
tacles (Fig. 3-23).
I (a) Standard coil. A helically wound coil of wire which forms a %ut” into
which coil bolts or rods (Sect. 3.4.1.4) may be threaded.
I (b) Tapped coil. The wire coils above may be tapped to accept standard
machine bolts. Such tapped coils will also accept coil bolts. Because of
I the dual thread, however, special care is needed in starting the coarse
thread coil bolt to prevent “cross-threading”.

I Cc) “Ferrule” or “Weld nut”. This is for use with standard bolts or rods with
standard threads. The nuts are of sufficient length and proper chemical
composition to be welded to the wires which form the anchoring
I element. Design strengths are shown in Table 3-6.

I
I 3.37
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
(a) Standard coil (b)Tapped coil (c) Weld nut (Ferrule)
I
Fig. 3-23 - Receptacles used in concrete inserts
I
3.38
I
I
I
I -
I

Fig. 3-24 - Bent wire and coil red inserts

Fig. 3-25 - Weage inserts (Not recommended for structural uses)

3.39
I
Table 3-6 - Design Strength of Machine Bolts in “Ferrules” or “Weld Nuts”
I
Bolt Dia.,
(in.1 Bolt Grade
Tensile
strength
Shear
strength Femle Data
I
(ASTM) P, (lb.1 V, (lb.1 Threads/in. Bolt Length
l/2
5t8
A307
A307
4820
7680
3330
5220
13
11
1
1-118
I
3t4 A307 11,360 7510 10 l-l/8
1 A307 20.600 13,350 8 l-l/4
I
3.4.2.2 Anchorage of inserts
I
Fig. 3-15 shows typical types of anchorage failures with inserts. The inserts
shown in Fig. 3-24 depend on bond with the concrete for anchorage. Those in Fig. 3-22
depend on mechanical anchorage and a shear cone type failure mechanism as discussed in
I
sect. 3.3.4.
3.4.2.3 Mechanical strength of inserts
I
If the insert is adequately anchored into the concrete, and the anchorage wires I
are properly welded to the receptacle, the strength of the insert is dependent on the
number and strength of wires or the connector capacity. It is most desirable to have the
bolt or the wires govern the connection strength, because they are more predictable and
I
ductile. Strengths of wires typically used in inserts are shown in Table 3-7.
3.4.2.4 Expansion inserts
I
Expansion inserts are devices placed into predrilled holes in hardened
concrete. The anchor develops a tension load-carrying capacity when expanding parts of
I
I
Table 3-7 - Capacity of Round Wire Used in Concrete Inserts
Yield
I
Leg Wire Wire strength,
Dia., in.
0.216
Grade
Cl008
lb
2000
I
0.223 Cl038 3900
0.225
0.240
0.260
Cl038
Cl008
Cl008
3700
2900
3550
I
0.281 Cl035 6000
0.306
0.340
0.375
Cl035
Cl035
Cl008
6900
7500
7450
I
0.440 Cl035 12.000

I
3.40
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I Fig. 3-26 - ‘Qpical expansion inserts

I
I 3.41
I
the insert exert lateral pressure against the sides of the hole. This is usually done by
I
tightening the connector bolt into the insert.
All expansion inserts ere proprietary. Examples are shown in Fig. 3-26. Each I
manufacturer has established the tensile and shear strengths of his devices by testing.
Typical ranges of tensile and shear strengths, taken from manufacturer’s catalogs are I
shown in Table 3-8. At the minimum recommended embedment depth, the tensile ca-
pacity agrees quite well with the “shear cone” concept described in Sect. 3.3.4
(Fig. 3-27). However, because of slip of the anchor in the hole, deeper embedment does
I
not proportionally increase the capacity. The upper limit of the capacity of the devices
is the strength of the connector bolt. Standard bolts (Sect. 3.4.1.1) are usually used as
I
connectors.
Edge distances for expansion inserts are more critical than for cast-in inserts. I
Expanding the insert in the direction of the edge should be avoided (Fig. 3-28). The
minimum edge distance should be at least five times the diameter of the connector bolt. I
The advantage of expansion inserts is that they can be placed in exactly the
right position after the precast members are in place.
measures when cast-in inserts are misplaced or left out.
They are often used as corrective
Proper performance of the
I
inserts is largely dependent on field workmanship. The holes must be drilled straight,
deep enough and the proper diameter. ‘lhe bolts must be tightened to sufficient torque,
I
sometimes requiring pneumatic impact wrenches. For these reasons, most manufacturers

Table 3-8 - Range of expansion bolt strengths


I
(From manufacturer’s catalogs)

Bolt Diameter Tensile Strength (lbs) Shear Strength (lbs) I


l/4

318
1500 - 3600

3200 -6000
1200-3500

2500 -8500
I
Ii2 4200 -11,600 7300 -15,200 I
518 7000 - 19,500 11,000 - 17,000

314 9200 -28,400 I 17,000 - 27,000 1


I
I
l-1/4
1 22,500

34.000
-33,000

-40,000
26,000

40,000
-39,000

- 63.000
I
(1) Strengths vary with concrete strength. Concrete strengths range from 3500 to
5500 psi.
I
(2) All values sre at minimum recommended embedment.
I
3.42
I
I
I
I
I
I
I
I
I
I
I Fig. 3-27 - Comparison of expansion bolt manufacturing
streqth with %heer cone” concept
catalog pullout

I recommend that the working strength of the insert not exceed one-fourth of the catalog
ultimate strength values.
I The performance of expansion inserts under stress reversals, vibrations or
earthquake is not sufficiently known. Further research on this aspect is recommended.
I
I
I
I
I
pig. 3-28 - Direction of expansion
I
3.43
I
I
3.4.3 Welded studs I
Stud welding is a semi-automatic process that can be used to weld fasteners of
many sizes and configurations. It is a fast, economical method of anchoring steel plates I
to precast concrete members. The process is described in Fig. 3-29. Most precast co-
ncrete manufacturing
Fig. 3-30.
plants have stud welding capability.
The most common stud configuration
Typical equipment is shown in
used in connections of precast concrete
I
are headed studs (Fig. 3-31) and deformed bar anchors (Fig. 3-32).

3.4.3.1 Headed studs


I
The strength of a headed stud may be limited
steel. The concrete strength is deter.mined as described in Sect. 3.3.4.
by either the concrete or the I
Steel specifications
D1.1(3.32)
and other aspects of stud welding are covered
. Headed studs are classified as either ‘shear connectors”
in AWS
or “other than shear
I
connectors”. ‘lhe minimum tensile strength, f,, of shear connector studs is 60,000 psi,
while the minimum tensile strength of others is 55,000 psi. For design purposes, the yield I
strength, fy, may be assumed to be 0.9fs(3’2). The maximum design tensile strength as
limited by the steel is: I
@Ps = $Abfy = 0.9 @%fs I
where $ = 1.0
Ab = area of the stud shank
I
to:
References 3.2 and 3.28 recommend that the shear strength of studs be limited
I
“Vs = 0.75 $Abfs (3-23)
I
I
I
The
stud
stud lip is placed against the work surface. When the trigger 18 puffed. the
is raised. A welding arc burns off surface contamination, melts the stud tip
I
and a small area on fhe work surface. The stud is forced into the molten area
and instantly welded to the surface.
I
F&.2-29-smdweldingproeeas
I
3.44
I
I
I
I
I
I
I
I
I Fig. 3-M - Typical shrd wektieg equipment

I Table 3-S gives standard dimensions of readily available headed studs. The
plate thickness to which the stud is welded should be at least 2/3 of the stud diameter.
I The quality of a stud weld may be evaluated by bending the stud to an angle of
30° from the original axis and striking it with a hammer. If the equipment is not proper-

I ly adjusted, the weld will fracture. This test is performed at the start of the work or
after any change in working conditions.

I
I
I
I
I
I
pip. 3-31- Applications of headed sh&
I
I 3.45
I
table 3-9 - Dimensions of headed studs, in. (mm)
I
I
I
Shank Heed Head Stock lengths’
(Nominal dimensions)
I
diameter diameter thickness
ds
l/4 (6)
dh

l/2 03)
th
3116 (5) 1, 2-l/2,4
(25, 64, 102)
I
3fa (10) 3/4 (19) 9132 (7) 3,4, 6
(76, 102, 152) I
l/Z (13) 1 (25) 5116 (8) 2, 3, 4, 5, 6, a

5/a (16) l-1/4 (321 5116 (a)


(51.16, 102, 127, 152, 203)

2,4, 6, a
(51,102,152, 203)
I
3/4 (19) l-114 (32) 3/a (10) 3, 3-l/2, 4, 4-112, 5, 6, 7. a
(76, 89, 102,114,127,152, 178, 2031 I
718 w2) l-3/8 (351 318 w0 3-l/2,4,5,6,7.6

* Other lengt
(89,102,127,152,1’/8,203)
I
I
I
I
I
I
I
I
Fig. 2-22 - AppllcJltions of deformed bar anchors
I
3.46
I
I
I 3.4.3.2 Deformed bar anchors

I Deformed bar anchors (Pig. 3-32) are anchored in the concrete by bond. The
deformations are indentations rather than projections as on reinforcing bars. While they
do not meet the ASTM specifications, tests have shown bond properties to be similar.
I Reference 3.2 suggests the following equation for development length of deformed bar
anchors:
I
I Ed - ";">'y
C
% 12 in. (3-24)

where:
I x
= as defined for Rq. 3-13a
= diameter of bar, in.
I db

fY
d 60,000 psi
for fy greater than 60,000 psi, the above value should be multiplied by:
I 60.000
2- f
I Y

As with reinforcing bars, when the deformed bar anchors are placed with more
than 12 in. of concrete below the bar (top bars), the above ~valuesshould be multiplied
I by 1.4.
Deformed bar anchors are available in diameters of l/4 in. to .5/S in., and
I lengths of 12 in. to 30 in.
3.4.4 Welding
I Loads are frequently transferred within a connection by welding. Welded con-
nections are efficient and performance depends on reliable workmanship and the compat-
I ibility of welding materials with the metals to be joined. Detailed discussion of welding
of connections in precast concrete is given in Ref. 3.3.
I Weld material is specified by an electrode classification number, e.g., E60, E70,
etc. The number represents the minimum ultimate tensile strength, F,, of the electrode,
I in ksi (e.g., E60 electrodes, F, = 60 ksi). The ultimate shearing strength of weld material
is assumed by the American Welding Society to be 0.60 Fu(3’33).

I Full penetration welds such as the V-groove welds shown in Figs. 3-33 and 3-34
transfer loads in direct tension or compression. If the electrodes recommended by the

I American Welding Society(3’32s 3*33) for various steels are used, and the welds are

I 3.47
I
properly made, full penetration welds will be stronger than the base metal, and can be so
I
assumed in design. Fillet welds and others which transfer loads through shear in the weld
must be designed for the actual applied loads.
I
3.4.4.1 Structural steel welding
Welding of steel plates, angles and other shapes should follow AWS D1.1(3*32).
I
Nearly all structural steel used in precast concrete connections is ASTM A-36. Thus it is
readily weldable with standard equipment.
I
Design procedures for structural steel welding have been established with steel
construction in mind, so are based on working stress methods. In precast concrete con-
I
nections, it is often more convenient to use factored loads and strength design proce-
dures. Two publications have suggested values to be used for welds in shear with I
strength (or “load factor”) design: the AWS Reinforcing Steel Welding Code(3’33) and the
AASHTO Bridge Specifications(3’34). In the former the nominal (ultimate) shear strength I
of welds to be used with factored loads is equal to the assumed ultimate shear strength,
0.6Fu. It implies that the capacity reduction factors, 4 , used in reinforced concrete
construction should be applied to this value. In the AASHTO Specification, the nominal
I
shear strength is specified as 0.45Fu, but does not use a $ -factor. Note that the
AASHTO Bridge Specifications allows only 0.2?Fu for working stress design, compared I
with 0.30Fu specified by AISC for building design(3.19). The comparisons are shown
below: I
Allowable
working stress
Allowable
design strength
I
AWS(3.33) 0.3Fu 0.6Fu x $I I
1.70 ( 4, = 0.85)
AASHTO(3.34) 0.27Fu 0.45Fu 1.67 I
For simplicity, it seems reasonable and conservative to assign a constant value
of 1.67 times the allowable working stress to the design strength, as is done in the
I
AASHTO Bridge Specifications. This is the basis of Table 3-10.
In welded connections, combined tensile and shear loadings are frequently
I
encountered. Design procedures using working stresses are given in the AISC Steel
Manual(3’35). By using the values in Table 3-10, these procedures can also be used for I
strength design.
I
3.48
I
I
I Various types of welds are shown in Fig. 3-33. The most commonly used are

I fillet welds and full penetration welds of either the vee-groove


Fillet welds are usually made assuming a 45’ fillet.
or bevelgroove type.
Table 3-11 gives strengths of fillet
welds using these assumptions.
I
Table 3-10 - Allowable working stress and design strength of welds(‘)

I Electrode Allowable
Working Stress (ksi)
Designc2)
Strength (ksi)

I I E60 18 I 30

I E80 24 40

I E90

El00
27

30
45

50

I Based on AISC Spec. for buildings. For bridges, use 90% of values.
Use factored loads and @ = 1 .O with these values.

I Table 3-11 - Strength of filet welds for building construction (1)

I
I
I
I
I
l/2 6.36 10.61 7.42 12.37
I 9/16 7.16 11.93 8.35 13.92

I 518

(1)
7.95 13.26

Use 90% of values for bridges. Assumes 45’ fillet.


9.28 15.47

(2) Use factored loads and c = 1.0 with these values.


I
I 3.49
I
I
I
I
I
I
I
I
Fig. 3-33 - Types of welds

3.4.4.2 Reinforcing bar welding


I
Welding of reinforcing bars is covered by AWS D1.4-79, “Structural Welding
Code - Reinforcing Steel” by the American Welding Society (3.33). Weldability is defined I
in that publication as a function of the chemical composition, as shown in the mill report,
by the following formula: I
C.E. =%C+ %++%++T+ %Ni %Cr
(3-25) I
where C.E. = carbon equivalent
For simplicity, some quality control engineers assume the last six terms of thii I
equation to be +0.05, resulting in the approximation:

C. E. =%C+%+ + 0.05 (3-25a)


I
AWS D1.4-‘79 indicates that most reinforcing bars can be welded. However, the
preheat and other quality control measures that are required for bars with high carbon
I
equivalents are very difficult to achieve. Fkcept for welding shops with proven quality
control procedures that meet AWS DlZ.l-79, it is recommended that carbon equivalents
I
be limited to 0.45% for #7 and larger bars, and 0.55% for #6 and smaller bars.
Most reinforcing bars which meet ASTM A615, Grade 60(3*36) will not meet the
I
above chemistry specifications. A615, Grade 40 bars may or may not be weldable.
I
3.50
I
I
I 45”. 600 45’ - 60”

I
I SingleV-groove weld
118”
Double-V-groove weld
I Full penetration welds

I
I
Fillet welds
I
-
t, = d,/5
d

I
I
I Flare-V-groove welds

I
I
I
I
I Flare-bevel-groove welds

I Fig. 3-34 - Typiiel reinfore* bar welds

I 3.51
I
?
Table 3-12 - Size of fillet weM required to develop full strength of bar I
I I 1
I
Bar perpendicular to plate,
welded one side
I
Plate fy = 36 ksi I
I

Bar
Gr. 40 bar, E 70 Electrode
Nominal weld Min Plate
Gr. 60 bar, E90 Electrode
Nommal weld Mm plate
I
size size in. I thickness, in. size, in. thickness, in.

3 3/16 l/4 3116 l/4


I
4 3/16 l/4 l/4 3fa I
I
I
I
I
ASTM Specification A706(3’37) provides for weldable reinforcing bars; however they are
not generally available. Mill reports should be checked for all bars to be welded.
I
Fig. 3-34 shows the most common welds used with reinforcing bars. Full pene-
tration groove welds can be considered the same as the nominal strength of the bar. The I
design strength of the other weld types can be calculated using the values from Table
3-10. The total design strength of the weld is: I
fw4vtw (3.26)

where
I
fW = unit design strength from Table 3-10
= length of weld (Fig. 3-34)
I
ft.4
tW
= thickness of weld (Fig. 3-34)
Tables 3-12 through 3-15 show welding required to develop the full strength of I
reinforcing bars.
I
3.52
I
I
I Table 3-13 - Siie of fillet weld required to develop full strength of bar

I Bar perpendicular to plate,


welded both sides

I Plate fy = 36 ksi

I
I I Gr. 40 bar, E 70 Elec~trode I Gr. 60 bar, E90 Electrode
BW Nominal weld Min Plate Nominal weld Min plate
size size in. thickness, in. size, in. thickness, in.
I 3 I/8 5/16 118 3/a

I 4
511/S
118
1
5/16
5116 I
l/8

3/16 I
318

l/2

I 6 1 3/16 1 7/16 1 3/16~ 1 l/2


7 1 S/16 1 7/16 I 3116 I 112
I 8 1 3/16 1 7116 I l/4 I 11/16

I 9

10
l/4
l/4
9/16
9/16
l/4
5116
11/16
718

I 11 5116 11/16 5/16 718

I 3.4.5 Reinforcing steel


Reinforcing steel used in connections meets ASTM A615(3’36) or
I ASTM A706(3*37). Load transfer between bars may be by welding (See Sect. 3.4.4.2), lap
splices or various types of couplers.
I 3.4.5.1 Lap splices
Requirements for bond development are discussed in Sect. 3.3.4. Lap splice
I requirements for reinforcing bars are given in AC1 318(3.11).
3.4.5.2 Couplers
I A variety of proprietary mechanical couplers are available for use in precast
concrete connections. Typical examples are shown in Fig. 3-35. Strengths of these
I
I 3.53
I
Table 3-14 - Design strength of connection with welded cross bar (kip)
I
I
I
I
I
Grade 40 Reinforcine bars: E’70 weld electrodes
I
” I

bl
12
\ #3

2.5
#4

3.3
#5
4.1
#6 #7 #a #9 #lO #ll
I
7.7
3.3 4.4 5.5
I
-
-. 4.1 5.5 6.9 ::: a.2
4.9 6.6 a.2 9.9 Il.2 Ia.Y
9.6 f 11.5 13.5 1 15.4 1 ;;:i 1 19.6 1
11.0
I
-It Grade 60 Reinforcing bars; E90 weld electrodes
I
I
#5 #6 #7 #a #9 #lO #11
I
5.3
7.1
6.4
a.5 9.9
I
a.8 10.6 12.4 14.1
10.6
12.4
14.1
12.7
14.8
17.0
14 .a
17.3
19.8
17.0
19.8
22.6
19.1
22.3
25.5
25.1
28.7 31.9
I
19.1 22.3 25.5 28.8 32.4 36.0
I
I
I
3.54
I
I
I Table 3-15 - Minimum length of weld to develop full strength of bar.
Weld parallel to bar length.
I
I
I
I Gr. 40 bar, E70 Electrode
Min Plate
Gr. 60 bar,

Minimum weld
E90 Electrode
Min plate
Bar Minimum weld
I size
3
length, in.
718
thickness, in.
3/16
length, in. thickness, in.
3116
4 l-1/8 3116 1-',,8 l/4
I 5
6
l-l/Z
l-314
/
:,;6 l-314
2f3,8
5/16
318
7 2 5/16 'i/16

I 8
9
10
2-114
2-5/a
2-718
3/8
7116
7116
2-518

3&
:,i/

9/16

I 11 3-l/4 l/2 3-314 518

devices are based on test data. Some are recommended for compression splices only,
I while others can be used for tension splices. Design should be based on the manufac-
turer’s recommendations. AC1 318(3*11) requires mechanical couplers to be capable of

I developing 125% of the yield,strength of the bars.


3.4.6 Dowels
I Reinforcing bars or structural steel rods are frequently used as dowels to con-
nect precast concrete components. These dowels may be cast in one member, and field
I placed and grouted into a preformed or predrilled hole in the other member, or they may
be field placed into both members. In most applications, these dowels are placed verti-

I cally and used only for alignment, or to resist horizontal loads from wind or earthquake.
Thus anchorage is normally not a problem.
Occasionally, a dowel will be required to resist tension. In this case, the bar
I must be sufficiently embedded to develop its design strength. However, the bond of the
grout to the concrete may be more critical. Tests(3.38) have shown that ordinary sand-
I cement grout in drilled holes is unreliable under direct tension loads. Anderson(3’3g)
reported on pull-out tests of dowels in holes which were preformed using a flexible
I
I 3.55
I
I
I
I
I
I
I
I
I
I
I
I
Fig. 3-35 - Proprietary mechanical couplers

metallic conduit (the type used for post-tensioning conduit). With bars up to #8, the full
I
strength of the bar was developed.
3.4.7 Post-tensioning steel
I
Post-tensioning is often used in connections when high ultimate tensile strength I
is desired, for example, in moment connections. Either ‘I-wire strands meeting ASTM
A416(3’40) or bars meeting ASTM A722(3*41) are used. The ultimate strength of strand is
either 250,000 or 270,000 psi. The ultimate strength of bars ranges from 145,000 to
I
160,000 psi.
In order to reliably measure the prestressing force, the tendon must be at least
I
15 or 20 ft long. Thus the prestressing feature of post-tensioned connections is normally
I
3.56
I
I
I only relied upon when long continuous tendons can be employed. Examples of this use
include vertical continuous post tensioning of columns and bearing walls and continuous
I frames in which the tendons also serve as all or part of the flexural reinforcement in the
beams. When shorter tendons are used, the primary value is in the high ultimate strength
I and ductility provided.
3.4.8 Pads ahd other bearing devices
I Bearing pads are used to distribute vertical loads over the bearing area. Some
types of pads also reduce force build-up at the connection by permitting small displace-
I ments and rotations. Their use is encouraged wherever applicable.
The performance of most bearing pads is a function of their deformation char-
I acteristics under service loads. Hence, these pads are designed using unfactored (ser-
vice) loads.
I There are a number of suitable materials and combinations of materials that
can be used for bearing pads. A few are described below with some design recommenda-

I tions. In some cases, various grades of bearing pads can satisfy these descriptions, but
exhibit widely different properties and behavior.
1. Commercial grade elastomeric (neoprene) pads are readily available and inex-
I pensive. However, these pads exhibit wide variations in shear deformation
characteristics and bearing strength, so they should not be used unless satisfac-
I 2.
tory performance test data is available.
Structural grade neoprene pads are those which meet the requirements of
I Section 25, Division 2 of the AASHTO Bridge Specifications (3.34). For optimum
economy, their use should be limited to places where uniform bearing is impor-

I tant, or when it is desired to reduce volume change restraints. Design recom-


mendations are shown in Fig. 3-36.
For high compressive stresses and/or large horizontal displacements,
I laminated pads consisting of layers of elastomer bonded between steel plates
can be used. Eech layer behaves in compression like an individual pad, but the
I shear deformation is a function of the thickness of the total assembly.
It should be cautioned that in unheated buildings, such as parking struc-
I tures, the unfactored bearing stress should be a minimum of 400 psi to keep the
pad from “walking” due to cyclical temperature conditions.

I 3. Laminated fabric bearing pads composed of multiple layers of 8-oz cotton duck
with a high quality natural rubber binder can sustain unfactored compressive
stresses up to 2000 psi. They are designed in a manner similar to elastomeric
I
3.57
I
I
pads except the shape factor need not be considered. These pads do not deform
I
as readily as elastomeric pads, so do not provide the same stress reducing
characteristics. The shear modulus, G, may be assumed to be 550 psi, unless I
more specific data is available.
4. Preformed pads composed of synthetic fibers and a rubber body are designed as
in (3) above, except that unfactored compression should be limited to 1500 psi.
I
The shear modulus can be assumed equal to 525 - 4v/3 unless more specific data
I
I
I
I
I
I
I
I
I
I
I
I
I
Fig. 3-36 - Design of structural gnde elastomeric beer- peds
I
3.58
I
I
I is available (v - unit unfactored shear stress). This type pad should meet the
requirement of Section 2.10.3 (L) of the AASHTO Bridge Specification (3.34).
I 5. Tetrafluorethylene (TFE) bearing pads reduce horizontal stresses because of
their low coefficients of friction. The unfactored bearing stress should not
I exceed 1000 psi when unreinforced or 2000 psi when reinforced with glass fibers
or similar material. TFE pads are sometimes used in combination with elast-

I omeric or fabric pads, or can be bonded to steel plates.


Higher compressive strengths are possible with some types of TFE pads,
provided the level of the bearing surface is held to very close tolerances. This
I is seldom practical in most precast concrete connnection applications.
6. A multipolymer plastic bearing strip is manufactured expressly for bearing
I purposes. It is a commonly used material for the bearing support of hollow-core
slabs, and is highly suitable for this application. The material has a compressive
I strength higher than the typical design range of concrete used in precast con-
struction.

I Tempered hardboard strips are also used with slabs to prevent concrete to
concrete bearing.

I 3.4.9 Cast-in-place concrete


Field-cast concrete is frequently used to transfer compressive forces in connec-
I tions, especially in ductile moment resistant frames, and in composite construction.
The rules of all cast-in-place reinforced concrete apply to field-cast concrete

I used in connections. Very often, connection areas will be heavily congested with rein-
forcement, so it is advisable to use small coarse aggregate, and to design concrete mixes
for relatively high slump.
I 3.4.10 Grout

I Many connections require the use of grout. Sometimes this grout is required
only for fire or corrosion protection, or for cosmetics. Other times, it is required to

I transfer compressive loads.


3.4.10.1 Sand-cement grout and dry-pack

I Most grout used in connections is a simple mixture of portland cement, sand,


and water. Proportions are usually one part cement to 2.5 to 3 parts sand. The amount

I of water depends on the method of placement.


Plowable grouts are highslump mixes used to fill voids that are either formed

I
3.59
I
I
in the field, or cast into the precast member. They are used within connections that are
I
heavily congested but not confined thus requiring some formwork. When such grouts are
used, the water-cement ratio is usually about 0.50. This is a relatively high value, result- I
ing in low strength and high shrinkage. These mortars also exhibit a tendency for the
solids to settle, leaving a layer of water on the top. Special ingredients or treatments I
can improve these characteristics, but add to the cost.
When very small spaces in confined areas are to be grouted, they are some-
times pumped or pressure injected. The confinement must be of sufficient strength to
I
resist the pressure. Rxamples are ducts for post-tensioning tendons or dowels. Some-
what less water is used than in flowable grouts, hence there is less shrinkage and higher
I
strength.
“Dry pack” is the common name used for very stiff sand-cement mixes. They I
are used when forming or other confinement is impractical, for example under column
base plates. Compaction is attained by hand tamping, using a rod or stick. I
3.4.10.2 Non-shrink grout
The shrinkage of sand-cement grout can be reduced-or compensated for-by
I
using proprietary non-shrink mixes, or by adding aluminum powder to the mix. Charac-
teristics and methods of testing non-shrink grouts are given by a Corps of Engineers I
specification(3’53).
Non-shrink grouts are designed to expand sufficiently during initial hardening I
and curing to offset subsequent shrinkage of the grout. Reference 3.53 classifies non-
shrink grouts by the method this expansion is accomplished.
1. Gas-liberating
I
2. Metal-oxidizing
3. Gypsum-forming
I
4. Expansive cement.
Some expansive ingredients may cause undesireable effects in some applica- I
tions, so manufacturer’s recommendations should be followed.
Proprietary grouts are prepared mixtures of cement, fine aggregate and the I
expansive ingredient. Aluminum powder added to ordinary sand-cement grout forms a
gas-liberating type of mixture. Extremely small amounts of aluminum powder are re-
quired - 50 to 60 millionths of the weight of cement used - about a teaspoonful per bag
I
of cement. ‘Ihe performance of the aluminum powder mixes are thus very sensitive and
not always predictable. Trial mixes should always be made with the particular cement
I
and sand that are to be used in the aluminum powder mixes.
I
3.60
I
I
I 3.4.10.3 Epoxy grouts

I Epoxy grouts are used when very high strength


to the concrete is necessary.
is desired, or positive bonding
They are mixtures of epoxy resins, as discussed in

I Sect. 3.4.11 and a filler material,


on the subject by Committee
usually sand. Reference 3.54 is a comprehensive report
503 of the American Concrete Institute.
The physical properties of epoxy compounds vary widely. The user should be
I familiar with the particular compound being used, either through experience or tests.
Test methods are given in Ref. 3.43. Of particular importance in some applications is
I the thermal expansion, which can be up to 7 times that of concrete.
Several railroads have used a different type of epoxy grout in shear keys on

I bridges. Instead of a pre-mixed


way, then pour in a low-viscosity
mortar, they place a well-graded
epoxy resin.
aggregate in the key-
This has the advantage of using a higher

I aggregate-resin ratio, thus it is more economical. It is also easier to mix, place and
clean up, and has a coefficient of thermal expansion that is more compatible, although
still about twice that of concrete.
I 3.4.11 Epoxy Resins

I Epoxy compounds are generally formulated in two or more parts. Part A is


most often the portion containing the epoxy resin and Part B is its hardener system.
Usually epoxy systems must be formulated to make them suitable for specific end uses.
I The epoxy compounds can be used to bond hardened concrete or other construction
materials to hardened concrete. They can also be used for grouting or pressure injection
I into cracks to restore the tensile strength of concrete or other materials. Epoxy resins
specially formulated to be moisture insensitive can be used to bond plastic concrete to

I hardened concrete or to repair concrete when water is present.


ASTM has a specification for Epoxy Bonding Systems (3.42) . The specification

I divides the systems into Types, Grades, and Classes.


hardened concrete and other materials to hardened concrete.
Type I is to be used for bonding
Type II is for use in bond-
ing freshly mixed concrete to hardened concrete.
I The term Grade is used to define viscosity. Grade 1 is low viscosity which
could be used to fill fine cracks. Grade 2 is medium viscosity and Grade 3 has a non-
I sagging consistency.
The Class determines setting time, which is affected by ambient
I temperature.
and 15.6’C).
Class A is to be used below 40°F (4.4’(Z), Class B between 40 and 60°F (4.4
Class C is intended for use above 60°F (15.6’C).

I
I 3.61
I
Epoxies are often suggested, but seldom used in precast concrete connections,
I
except for grouting anchor bolts or dowels into pre-drilled holes. They have also been
used for repair or modification of connections in the field. I
Several types of connections suggest the application of epoxy bonding com-
pounds. For example, a precast corbel could be attached to a precast column. The I
strength and ductility are questionable, however, and additional research is suggested.
3.5 Special Designs I
The special connection elements described in this section have become stan-
dard in the industry. They usually involve combinations of load transfer mechanisms and I
devices. ‘Ibe design procedures have been developed from tests and analysis.
3.5.1 Reinforced concrete corbels I
One of the most common elements in beam-to+olumn connections is the
reinforced concrete corbel (sometimes called a bracket” or “haunch% Corbels have I
been extensively researched(3.43-3.451. The first major research work was done by Kriz
and Raths at the Portland Cement Association (3s43). A total of 195 corbels were tested, I
of which 124 were subjected to vertical load only and 71 to combined vertical and hori-
zontal load. This research resulted in the empirical design equation: I
(l/3 + 0.4Nu/Vu’
wn = 6.5bd 2/T (1 - 0.5d’a) (‘Ooo p)o.8N ,v
10 u” I
(3.27) I
This equation appeared in the PC1 Manual for Design of Connections (3.11 and in
the fist edition of the PC1 Design Handbook, along with design aids to assist in solving
I
it.
Equation 3-27 was approximated and simplified for inclusion in the 1971 ver- I
sion of the AC1 Building Code as Rq. 3-28 and was retained in the 1977 Code(“%

I$ V, = $ 5.1mu] [l - 0.5 a/d]


I
X + 160 dg) +d (3-28)
I
In Rqs. 3-27 and 3-28:
I
$ = 0.85
b = width of the corbel I
P = (4 + A,,h)/bd,not to exceed 0.013
I
3.62
I
I
I Other parameters are as shown in Fig. 3-37.

I The other major research on reinforced concrete corbels was done at the
University of Washington under the direction of Prof. Alan II. Mattock(3.44s3-45). Tnis
study included the testing of 28 corbel specimens, plus a review of the Kriz-Raths data.
I Mattock showed that the strength of a reinforced concrete corbel could be safely pre-
dicted by considering @-seFig. 3-37):
I (1) a moment equal to Vua + N,(h - d)
(2) a simultaneous horizontal tensile force, N,
I (3) shear resisted by shear friction as discussed in Sect. 3.3.2.3.
This approach resulted in the design equations suggested in the second edition
I of the PC1 Design Handbook(3’z). Referring to Fig. 3-3’7, the main tensile reinforcement,
As + An, is the greater of that determined by Rqs. 3-29 and 3-30:
I (3-29)

I (3-30)

I
I
I
I
I
I
I
I
concrete corbels
I Fig. 3-37 - Dfsign of

I 3.63
I
vu I
A (3-31)
vh = q

In these equations:
I
$
fy
=
=
0.85
yield strength of As + An
I
fyv
‘k
=
=
yield strength of Avh
effective shear-friction coefficient, as calculated in I
Sect. 3.3.2.3
vu = applied factored vertical force, limited to values shown in I
Table 3-l.
A review of reference 3.44 indicates that Bq. 3-31 misinterpreted the Mattock
and Kriz-Raths reports. The recommendations were that Avh should be equal to 0.5 A,,
I
regardless of whether the moment or shear-friction steel requirements are critical.
Thus, when used with Eqs. 3-29 and 3-30, a more proper equation would be:
I
Avh = 0.5
1
‘As + A,,’ - Nu/f Y (3-31a) I
This steel is provided to prevent a premature diagonal tension failure. I
Reference 3.2 also requires that:
(11 The minimum reinforcement requirements of Sect. 3.3.2.3 should be met.
(21 The horizontal reinforcement, Avh, should be distributed over a depth of
I
d/2 from the flexural reinforcement, A,.
(31 All reinforcement should be adequately anchored by welding, hooks, or
I
development length.
The above design approach is being considered for inclusion in future versions I
of the AC1 Building Code.
3.5.2 Structural steel haunches
I
Structural steel is often embedded in a precast concrete column or wall panel
to serve as a haunch or corbel. Numerous examples are shown in Part 4. They are often
I
easier and less expensive to fabricate than reinforced concrete corbels, and can be
designed to be stronger and more ductile. I
Design procedures are given in the PC1 Design IIandbook(3’2), based on a paper
by Raths(3’46). The method is a conservative approach, based on statical analysis, with I
the approximate assumptions shown in Fig. 3-38.
I
3.64
I
I
I
I
I
I
I Fig. 3-38 - Assumptions used in design of embedded
steel haunch by Reference 3.2

I The design strength of the section is:

0.85f; bke
(3-32)
I “c = 3.67 + 4a/Re

v, = nominal strength of the section controlled by concrete, lb


I a
Le
=
=
shear span, in.
embedments depth, in.

I b
The effective
= effective width of the compression block (See Fig. 3-39)
width of the compression block for double flanged members are
assumed as shown in Fig. 3-39(a), provided steps are taken to assure good compaction of
I
I -?-l-
1 \
z
- Reinforcement
welded to tube

I I--i
L
‘b=2w’

I
I
I
I
Fig. 3-39 - Effective width of embedded
I
I 3.65
I
the concrete and/or confinement of the concrete around the section. Holes in the em-
I
bedded part of the member (more than 1 in. diameter) aid compaction.
It should be noted that there must be adequate concrete and/or superimposed I
axial dead load above and below the haunch to develop the compressive forces indicated
in Fig. 3-39 in order for Eq. 3-32 to be valid. If not, it may be possible to develop the I
force couple by using reinforcing bars in tension, in a manner similar to that shown
below. I
Additional capacity can be obtained by welding vertical reinforcing bars to the
steel section, as shown in Fig. 3-39(b). The additional capacity is calculated by conserva-
tively assuming that the reinforcement acts at the centers of compression, and that
I
reinforcement nearest to the applied load is balanced by reinforcement near the end of
the steel member. Thus:
I
3 Asf
v, = 3.67 + 4a/g,, (3-33) I
As (12a + 2k,)
I
A;; = (3-34)
12a + llf,,

where:
I
vr = additional nominal strength of the section provided by rein-
forcement
I
A, = area of vertical reinforcement nearest to the applied load
(assumed to be located at a,/6 from face) I
A; = area of vertical reinforcement near the end of the steel
section I
(assumed to be located at 11 9.,/12 from face)
fy = yield strength of the reinforcement I
The total design strength of the section is thus:
lpn = L$‘vc + Vr’ (3-35)
I
r$ = 0.85 I
The design strength of the steel section can be determined by:
Flexural design strength: I
I
3.66
I
I
I
0” 2% (3-36)
I n a

Shear design capacity:

I +V, = @CO.55fy h w)
where:
(3-37)

I z,
fy
=
=
plastic section modulus of the steel section (see Table 3-16)
yield strength of the steel

I h,w
$ =
= depth and thickness of steel web, respectively
0.90

I Note: Plastic design criteria for structural steel does not require the use of
a +-factor . However, the load factors used are 1.7 (D + L). Therefore, when using
I steel plastic design with concrete load factors (1.4 D + 1.7 L), the use of $ = 0.90 is
recommended by the PC1 Connections Committee in order to provide approximately the
I same overall factor of safety.
For steel shapes projecting equally from each side of the member, with ap-

I proximately symmetrical loading, the design strength on each side as governed by the
capacity of the concrete can be calculated by:

I 0.85 ff bt,
bv, = 4 3 (3-38)

I Horizontal forces, Nut are resisted by bond on the perimeter of the section. If
the bond stress resulting from factored loads exceeds 250 psi, headed studs or reinforcing
I bars can be welded to the section.
In a study by Marcakis and Mitchell(3’47) the approximations of Fig. 3-38 were

I refined using the accepted strain and stress distribution assumptions in Fig. 3-40. The
resulting equations, listed below, were verified by testing 25 specimens.

I 0.85 f; b’l,
‘c = 1 + 3.6e/!le
(3-39)

I For the additional contribution of reinforcement


shape, and with A; = As :
welded to the embedded

I
I
I 3.67
Table 3-16 - Plastic section moduli and shape factors

1.12 Lapprox)

E 1.70
6

th’ fort < h 1.27 fort < h

bh'
h 12 2

@ b

2As f
(3-40)
6e/xe
’ + 4.8s/Le - 1

As in Eq. 3-35, $v, = $(vc + vr)

3.68
I
I
I
I
I
I IdPure sll*ar (b)Pura

pig. 3-40 - Strain and stress distributions


Yomanl (clGenerrl Loadln~

assumed in Ref. 3.47

I Notation for the above equations are described in Fig. 3-41.


Marcakis and Mitchell also recommended the following, based on their tests:

I (1) In a column with closely spaced ties above and below the haunch, the
effective width, b, can be assumed as the width of the confined region, or
2.5 times the width of the steel section, whichever is less.
I (2) For thin-walled members, such as the tube shown in Fig. 3-41, the inside
should be filled with concrete to prevent local buckling.
I (3) When the supplemental reinforcement, As and A; is anchored both above
and below the members, as in Fig. 3-41, it can be counted twice.

I (4) The critical section for bending of the steel member is located a distance
V,,/(O.SS f;? b) in from the face of the column.

I
I
I
I
I
I
Fig. 3-41 - Assumptions used in design of embedded steel haunch by Ref. 3.47
I
I 3.69
I
It can be seen that if the steel section projects from both sides, with equal
I
load on each side as in Fig. 3-40(a), Eqs. 3-39 and 3-40 are still applicable, with ae =
column width and e = 0. Because of the chance of unbalanced loads, a minimum eccen- I
tricity should be used. If a minimum value of e/!Ze = 0.5 is used, Eq. 3-39 becomes very
close to the PC1equation (Fq. 3-38). I
Further comparison of the PC1 Handbook equations (Eqs. 3-32 through 3-34)
with the comparable Marcakis and Mitchell recommendations @.q. 3-39 and 3-40) show
reasonable agreement when similar assumptions of embedment and reinforcement place-
I
ment are used. The PC1 equations are generally more conservative.
Using recommendation (4) above, the flexural design of the steel section would
I
be:
$Zsf
I
$‘n = a + VU’( O.i5f;bI

Another study on this subject at the University of Washington(3.55) confirmed


I
the conservative nature of the PC1 Design Handbook equations. However, that research
indicated that assuming the development of full compression under both flanges of a
I
wide-flange section, as implied in Fig. 3-39(a) is not correct.
With the benefit of these two research reports (Ref. 3.47 and 3.55) more I
accurate recommendations will undoubtedly be made in the near future.
3.5.3 Dapped-end beams
I
In order to keep the depth of floor or roof construction to a minimum, it is
frequently desirable to recess, or “dap” the end of a member. This can result in severe
I
stress concentrations at the re-entrant corner. When a connection utilizes such a dap-
ped-end, several possible failure modes must be investigated, as discussed in I
Section 3.2.2.
Design equations for the reinforcement are given in the PC1 Design I
Handbook(3’2). Several of these equations resulted from tests performed by C. II. Raths
of Raths, Raths, and Johnson, Willowbrook, IL. The results of these tests are unpublish-
ed, but are discussed briefly in Reference 3.8.
I
More recently, a comprehensive investigation was conducted at the University
of Washington. The results were reported in a Masters Degree Thesis(3.48) and summar-
I
ized in Reference 3.9. This program in general verified the approach of the PC1 Design
Handbook, but recommended some modification. I
The most common method of reinforcing a dapped-end member is as shown
I
3.70
I
I
I
I
I
I
I
I (a) Schematic of kinforcement Force resist@ system for vertical reaction

I Fig. 3-42 - Dapped-end beam design

I schematically in Fig. 3-42(a). The forces for which this reinforcement is designed can be
determined by considering the simple truss analogy shown in Fig. 3-42(b), plus reinforcing
I for the axial tension, N,. It can be seen that the most critical reinforcement is the
horizontal bars A, + An, and the vertical ‘hanger” bars Ash. However, the tests showed

I that when these two possible failure modes are adequately covered, other modes develop
as discussed in Section 3.2.2.

I The following equations are from the PC1 Design Handbook, modified as indi-
cated by more recent data. The failure modes investigated are as shown in Fig. 3-7.

I 3.5.3.1 Moment and Axial Tension


From Fig. 3-42(b), the tension in the primary horizontal reinforcement,

I As + An is VJtano where tana = j,,/a . this is equivalent to calculating the canti-


lever moment of Vua, and resisting it by the internal moment couple. In addition, this

I reinforcement must resist the direct axial tension and the small moment caused by the
horizontal reinforcement being placed eccentrically to the applied axial force, N,. Thus:

I As =
Vua + Nu (h - d)
Qfy j,,d
(3-42)

I NU
(3-43)
An = r
Y
I
I 3.71
I
where:
I
4
a
=
=
0.90
shear span, in. I
h = depth of the member above the dap, in.
d = distance from top to center of the reinforcement, A,, in. I
fy = yield strength of the flexural reinforcement, psi
JU
= distance from the centroid of the compression block to center
of (As + A,,,
I
The PC1 Design Handbook
equating $ ju = 0.85 in order to simplify calculations.
uses a conservative approximation
When this is done, Rq. 3-42 and 3-
of
I
43 can be combined and simplified to:
I
(3-44)
I
The Handbook illustrations show the shear span, a, measured from the point of
load to the end of the deep portion
measured to the centroid
of the beam.
of the hanger reinforcement,
Reference 3-9 required that it be
Ash. Examination of Fig. 3-42(b)
I
shows the latter
condition
assumption to be correct. (Note: The Handbook approximated
by suggesting the load be assumed at 3/4 of the extended end, and specified the
this
I
hanger reinforcement
beam.)
start at a maximum of l-l/Z in. from the end of the full depth
I
3.5.3.2 Tensile Cracking at the Reentrant Corner

A tensile crack will emanate from the reentrant corner of a dapped-end beam
I
at a very low load. The Raths test, some of which were conducted on unreinforced
members, showed that this crack starts at a principal tensile stress of about 4fic for
I
normal weight concrete. Once the crack starts, the failure is very sudden and brittle in
an unreinforced member. The limited number of sand-lightweight specimens tested by I
Raths failed at a principal tensile stress of about 2.75 .&, and were less predictable.
The principal tensile stress is calculated by: I
(3-45) I
where:
fp = principal tensile stress, psi
I
I
3.72
I
I
I ft
= flexural tensile stress plus axial tension

I = Va + M/2
bh2/6
N 6Va
+i3i=b7;z+ni
4N

I v =
=
vertical load, lb
N axial load, lb
V
I f,
a
=
=
shear stress = m, psi
shear span, measured from point of load to reentrant corner,

I b =
in.
member width, in.
h = depth of the extended end, in.
I Because of the brittleness of the failure the unpredictability of other para-
meters (such as load placement and magnitude of axial load), and the possibility of cracks
I developing during handling, the concrete tensile strength should not be relied upon in
narrow members such as beams and joists.
I Thus in dapped-end beams the hanger reinforcement, Ash, must take the total
vertical shear:

I Ash =TvU (3-46)

I where
$ = 0.85
I vu = applied factored load, lb
Ash = vertical bars across potential re-entrant corner crack, sq in.
I fy = yield strength of Ash, psi
3.5.3.3 Direct Shear at the Extended End
I This reinforcement is determined by shear friction. The requirements are not
additive to the flexure requirements above. The PC1 Design Handbook recommends that
I 2/3 of the shear-friction steel requirements be supplied at the bottom of the extended
end (As in Fig. 3-42a), and l/3 be distributed in the bottom 2/3 of the extended end (AVh
I in Fig. 3-42a). The design equations then become:

2vu Nu (3-47)
I ‘s’ An = 3q +q

(but not less than calculated by Eq. 3-44)


I
I 3.73
I
I
Avh v” (3-48)
I
where
+
fy
=
=
0.85 I
yield strength of reinforcement, psi
I
u = 1000 A2 bh u
e v” I
(See Eq. 3-13a for definition of A )
The recommended minimum reinforcement requirements are: I
I
I
unless one-third more than that required by either Rq. 3-47 or 3-48 is provided.
The shear friction requirements are an extension of the recommendations for
I
corbels. ‘Ihe comments following Rq. 3-31 are also applicable here. IQ. 3-31(a) is prob-
ably more proper than Rq. 3.48. The Raths and Mattock-Chan tests incorporated rein- I
forcement calculated in this manner and direct shear was not critical.
The PC1 Design Handbook limits the shear stress, Vu/$bd, to 8 X 4. This I
is based on Sect. 11.5.6.8 of AC1 318-77(3*111,but was probably a misinterpretation. A
more appropriate upper limit would be as specified for shear-friction design in
Sect. 11.7.4 of AC1 318-77, i.e., 0.2 f; or 800 psi.
I
3.5.3.4 Diagonal Tension in the Extended Rnd I
In the Raths tests, the final failure of all of the reinforced specimens was a
result of compression in the extended end. The failure started as diagonal tension crack- I
ing, which eventually outlined an arch between the reaction and the applied load. In
several of the specimens, this failure occurred before the primary horizontal reinforce-
ment (As) or the hanger bars (Ash) yielded. This led to the recommendations in the PC1
I
Design Handbook for both vertical and horizontal reinforcement in the extended end as
follows (see Fig. 3-42a):
I
I
3.74
I
I
I fy+2 xbdfi;
1
I
(3-52)
I
The Mattock-Chan tests showed cracking which indicated this as a potential
I failure mode, but the premature compression failure did not occur. An examination of
the reinforcement details of both test programs, Fig. 3-43, may show why. In the

I Mattock-Chan specimens, the horizontal bars had more positive anchorage and were
better distributed. They also had perimeter “framing bars” which were not included in
the design. In the Raths specimens, the low placement of the Avh bars and the load
I placement permitted the cracks to bypass the bars and proceed uninterrupted to the load
point. While the Mattock-Chan tests did not indicate a need for the vertical reinforce-
I ment, their report recommended such reinforcement be used if the ratio of a/d exceeds
one.
I 3.5.3.5 Diagonal Cracking in the Undapped Portion
In the Mattock-Ghan tests, another possible failure mode was observed, diago-
I nal cracking in the lower part of the undapped portion. These cracks assumed an angle of
approximately 45O. To reinforce against this possible failure mode, the report recom-
I mends that the primary horizontal reinforcement be extended a full development length
beyond this potential crack (see Fig. 3-42a).
I 3.5.3.6 Bearing in the Extended End
As in all connections, the bearing should be checked as discussed in
I Section 3.3.1. Normally, the reinforcement required for the other failure modes will be
adequate to confine the bearing area.
I 3.5.3.7 Detailing Considerations
Anchorage of reinforcement in dapped-end connections is very important.
I Examination of the Mattock-Chan and the Raths tests suggest the following detailing
recommendations (see Fig. 3-44).
I 1. ‘lhe main reinforcement, As + An, should be positively anchored at the end
of the beam. Welding to a plate or confinement angle at the end of the
I beam is recommended.
2. ‘lhe hanger bars, Ash, should be placed as close to the end of the full-

I
I 3.75
I
I
I
-z-4+3 ClW.d sllrrupl
I
I
I
I
I
I
2-n3,3-t3 or Z-*0

I
I
I
0 Mattock-than teata

Fig. 3-43 - Specimens for dapped+nd beam tests I


depth portion of the beam as cover requirements permit. These bars
should be closed ties which wrap around horizontal reinforcement in the I
top and bottom of the beam. When multiple bars are required, “bundling”
is suggested as a means to place the center of resistance as near the end I
as possible.
3. ‘lke horizontal bars, Avh, should be anchored near the end of the beam.
Vertical bars should be placed at the end, and the horizontal bars wrapped
I
around them. At the other end, these bars should extend a minimum
of 1.7 ad past the dap. These bars should be distributed throughout at
I
least 2/3 of the depth of the extended end.
I
3.76
I
I
I
I
I H

I
I
I
I Fig. 3-44 - Reinforcement for Dapped-lhd

I 4. The horizontal bars, As + A,, should extend into the beam a distance equal
to (H -d + kd ) past the end of the beam, where H is the depth of the

I undapped portion. This assures that the bars will be developed beyond a
45’ crack which starts from the bottom corner of the undapped portion.
Note that in most applications, the As + An and the Avh bars will be more
I than 12 in. above the bottom of the beam. Thus they are classified as “top
bars” by the AC1 Code13’11), and increased development length,Ild , is
I required.
In addition to the above, the PC1 Design Handbook recommends that the depth
I of the extended end should be at least one-half the full depth of the member. If a great-
er dap is required, hanger connections of the type discussed in Section 3.5.4 are recom-

I mended.
3.5.3.8 Alternate Placement of Reinforcement
I In some applications, the primary reinforcement is placed diagonally as shown
schematically in Fig. 45(a). In this case the forces are resolved as shown in Fig. 45(b).
I The reinforcement requirements are thus:

VuLF-Z
I Ash = I# fy“U cosa z
4 fy d

I (3-54)

I
3.77
I
I
I
I
I
I
I
(a) sehemetic of reinforcement 0 Force resisti system for verticel reaction
I
Fig. 3-45 - Alternate design for dapped-ends

The shear-friction (Sect. 3.5.3.3) and bearing requirements (Sect. 3.5.3.6) must
I
also be satisfied.
Anchorage of the bars is, again, very important, It is sometimes difficult to
I
get proper anchorage of the diagonal bars, Ash, especially in the extended end. Careful
consideration of bend radii, hook diameters and other details is necessary. Welding to I
cross bars or confinement angles may be required.
3.5.4 Hanger Connections I
Hangers are similar to dapped ends, except that the extended, or bearing end is
steel instead of concrete. They are used when it is desired to keep the structural depth
I
very shallow. Hangers are exceptionally stable during erection.
3.5.4.1 Cazaly Hangers
I
The basic Cazaly hanger, developed by Lawrence Cazaly of Ontario, Canada,
has three basic components (Fig. 3-46ah
I
1. A cantilevered bar which supports the unit containing the hanger. The bar
is kept in equilibrium by pressure from the concrete at its interior end and I
by tension in the strap.
2. A strap which transfers the vertical load to the bottom of the unit. (In I
prestressed members it also serves as anchorage zone reinforcement.)
3. Dowels.
Design assumptions and recommendations for Cazaly hangers are given in the
I
I
3.16
I
I
I
I
I
I (a) Basic components cb) Designasumptions
Fig.3-46-cazalyhanger
I Canadian Prestressed Concrete Institute (CPCI) Handbook(3’4g) as follows (see
Pip. 3-46b).
I 1. The cantilevered bar is usually proportioned so that the interior reaction
from the concrete is 0.33 Vu. The hanger strap should then be proportion-
I ed to yield under a tension of 1.33 Vu

I (3-55)
where:
I fy
4
=
=
yield strength of the strap material
0.90

I When the strap yields, it applies a uniform load to the bar. The point of
zero shear is then 0.75s from the front face.
3. Vu may be assumed to be applied 0.5 in. from the face of the seat. The
I remaining part of the moment arm is the width of the joint, g. It is there-
fore important that the joint width used in analysis is not exceeded in the
I field.
4. ‘lhe moment in the cantilevered bar is then given by:
I I$ = Vu (0.5 + g + 0.375s) = 0 fy bd2/6 (3-56)
where
I fy = yield strength of the bar material
0.90
0 =
I Other notation as shown in Fig. 3-46b.
If the bar is proportioned to take this moment at the yield stress, but

I using elastic section properties, the shear and tensile forces can usually be
neglected.
5. ‘lhe bearing pressure creating the interior reaction may be assumed to be
I
I 3.79
I
T,. The bearing length Qb is then given by:
I
I
The exterior cantilever should have a minimum length of (g + 1) in. Most I
hangers in practice have cantilever lengths of 2-l/2 to 3-l/2 in.
6. TO maintain the conditions
must have a length:
of equilibrium assumed, the interior cantilever I
7.
(1.5 + 3g + s + O.!iQb) in.
The minimum total length of bar is then:
I
8.
(2.5 + 4g + 2s + 0.5Qb) in.
The weld connecting the strap to the cantilevered bar should be designed I
to a higher factor of safety than the rest of the strap. A weld throat
stress (at the hanger design load) of 20 ksi, based on the gross length of I
weld (ignoring end effects) has proved satisfactory.
9. Longitudinal dowels, An, should be welded to the cantilevered
transmit the axial force, NU:
bar to I
An = q
NU
(3-57)
I
where I
fy = yield strength of the dowel
$ = 0.90 I
In a series of 52 tests performed at the University of Toronto (3.50) , it was
found that nearly all failures were a result of shearing of the concrete at the inside edge
of the strap (see Fig. 3-46b). A review of the test data indicates that the lower dowel
I
and the area confined witbin~the strap can be conservatively proportioned using effective
shear-friction described in Sect. 3.3.2.3:
I
I
where I
$ = 0.85
fy = yield strength of lower dowels, psi I
I
3.80
I
I
I 1400 X2 bh
ue= v
I U

Vu (mu) = 0.30 A2 f; bh SC000 X2 bh (3-59)


I 3.5.4.2 Loov Hanger
Analytical and experimental studies reported by Loov(3.51) suggest a less
I complex and more ductile hanger, illustrated in Fig. 3-47. This connection is designed
using the following equation:
I
I
I
I
I
I
I
I (a) Eiasic components

I
I
I
I
I
I 3.81
I
I
Ash v” (3-60)
= Bfys cosa
I
where
@
fys
=
=
0.85
yield strength of Ash
I
N”
A, = “f;l (I + dh--a$ (3-61)
I
where
I
4 = 0.90
fy = yield strength of An I
The steel bar is proportioned so that the bearing strength of the concrete is
not exceeded, and to provide sufficient weld length to develop the diagonal bars. Bearing I
strength is discussed in Sect. 3.3.1. thus, from Eqs. 3-5 and 3-8, and referring to
Fig. 3-47a: I
fbu = 4 45 x
(3-62) I
where
$ = 0.70
I
x = 1.0 for normal weight concrete
= 0.85 for sand-lightweight concrete I
= 0.75 for all lightweight concrete
b2 = member width I
bl = width of the steel bar
The connection should be detailed so that the reaction, the center of compres-
sion and the center of the diagonal bars meet at a common point, as shown in Fig. 3-47.
I
The compressive force, C, is assumed to act at a distance a/2 from the top of the bearing
plate. Thus:
I
(3-63) I
where
Nu (h - d)
(3-64)
I
C = Vu tsna + d - a/2
I
3.82
I
I
I For most designs, the horizontal bars, An, are placed very close to the plane of

I bearing. Thus the term (h - d) can be assumed equal to zero, simplifying Pqs. 3-61 and 3-
64.
In the Loov tests(3’51) specimens in which beam shear was designed in accord-
I ance with ACI-318(3’1’), shear failure occurred before full capacity of the section was
reached and at loads significantly below the calculated ultimate. Thus, Loov recom-
I mended that stirrups in the beam end be designed to carry the total shear.
3.5.4.3 Variations
I Several variations of the Cazaly and Loov hangers have been used
successfully. Some of these variations are shown in Part 4.
I
I
I
I
I
I
I
I
I
I
I
I
I 3.83
I
RRFRRRNCES - PART 3
I
3.1 “PC1 Manual on Design of Connections for Precast, Prestressed Concrete”, First
I
3.2
Edition, 1973, Prestressed Concrete Institute, Chicago, IL.

“PC1 Design Handbook, Precast, Prestressed Concrete”, Second Edition, 1978,


I
Prestressed Concrete Institute, Chicago, IL.

3.3 “PC1 Manual for Structural Design of Architectural Precast Concrete”, First I
Edition, 1977. Prestressed Concrete Institute, Chicago, IL.
3.4 “Connection Details for Precast, Prestressed Concrete” by the Australian Pre-
stressed Concrete Group. Published by Cement and Concrete Association of
I
Australia about 1965.

3.5 “Structural Design Manual-Precast Concrete Connection Details”, Society for


I
Studies on the use of Precast Concrete, Netherlands. Published by Beton-Verlag
GmbH, Dusseldorf, 1978.
I
3.6 Phillips, W. R. and Sheppard, D. A., “Plant Cast Precast and Prestressed Concrete
- A Design Guide”, the Prestressed Concrete Manufacturers Association of Cali-
fornia, Inc., 1980. Available from the Prestressed Concrete Institute, Chicago. I
3.7 Speyer, Irwin J., “Considerations for the Design of Precast Concrete Bearing WalI
Buildings to Withstand Abnormal Loads”, PC1 Journal, v. 21, no. 2, March-April,
1976. I
3.8 Martin, L. D., “Background and Discussion on PC1 Design Handbook Second Fdi-
tion”, PC1 Journal, v. 25, no. 1, Jan-Feb, 1980. I
3.9 Mattock, A. H., and Chan, T. C., “Design and Behavior of Dapped-End Beams”,
PC1 Journal, v. 24, no. 6, Nov-Dee, 1979. I
3.10 “Expansion Joints in Buildings ‘I, Technical Report No. 65, prepared by the Standing
Committee on Structural Engineering of the Federal Construction Council, Build-
ing Research Advisory Board, Division of Engineering, National Research Council, I
National Academy of Sciences, 1974.
3.11 “Building Code Requirements for Reinforced
Concrete Institute, Detroit, MI.
Concrete (AC1 318-77)“, American I
3.12 Kriz, L. B., and Raths, C. H., “Connections in Precast Concrete Structures -
Bearing Strength of Column Heads”, PC1 Journal, v. 8, no. 6, Dee, 1963. Also
I
Portland Cement Association Research and Development Laboratories

3.13
Bulletin D73.

Williams, A., “The Bearing Capacity of Concrete Loaded over a Limited Area”,
I
3.14
Technical Report 526, (Aug, 19791, Cement and Concrete Association, London.

ACI-ASCE Committee 326 (now 426), “Shear and Diagonal Tension”, AC1 Journal,
I
v. 59, no. 1, Jan, 1962, and no. 3, Mar, 1962.
I
3.84
I
I
I 3.15 AC1 ASCE Committee 426, “‘lhe Shear Strength of Reinforced Concrete
Members”, Chapter 1-4, Proceedings, ASCE, v. 99, no. ST6, June, 1973; Chapter 5,
I Proceedings, ASCE, v. 100, no. ST8, August, 1974.

3.16 “Shear in Reinforced Concrete”, vols. 1 and 2, AC1 Publication SP-42, American
I 3.17
Concrete Institute, Detroit, MI.

AC1 Committee 318, “Commentary on Building Code Requirements for Reinforced


Concrete (AC1 318-63)“, American Concrete Institute Publication SP-10, 1965.
I 3.18 British Standards Institute, “Code of Practice for the Structural Use of Concrete”,
CPllO, November, 1972.
I 3.19 “Specification for the Design, Fabrication and Erection of Structural Steel for
Buildings”, American Institute of Steel Construction, effective Nov 1, 1978.

I 3.20 Commentary on Ref. 3.19.

3.21 Mast, R. F., “Auxiliary Reinforcement in Concrete Connections”, Journal of the


I Structural Division, American Society of Civil Engineers, v. 94, ST6, June, 1968.

3.22 Mattock, A. H., ‘Shear Transfer in Concrete Having Reinforcement at an Angle to


I the Shear Plane”, Publication SP-42, Shear in Reinforced Concrete, American
Concrete Institute, Detroit, MI, 1974.

I 3.23 Raths, C. H. and Mattock,


no. 2, March-April, 1977.
A. H., Discussion of Reference 3.44, PC1 Journal, v. 22,

3.24 Birkeland, H. W., Class Notes for Course on “Precast and Prestressed Concrete”,
I University of British Columbia, Spring, 1968.

3.25 Shaikh, A. F., “Proposed Revisions to Shear-Friction Provisions”, PC1 Journal,

I 3.26
v. 23, no. 2, March-April, 1978.

Mattock, A. H. and Martin, L. D., Discussion of Ref. 3.8, PC1 Journal, v. 25, no. 6,
Nov-Dee, 1980.
I 3.27 “Superior Precast Concrete Handbook”, Superior Concrete Accessories, Inc., San
Diego, CA.
I 3.28 “KSM Structural Engineering Aspects of Headed Concrete Anchors and Deformed
Bar Anchors in the Concrete Construction Industry”, KSM Fastening Systems

I 3.29
Division of Omark Industries, Moorestown, NJ.

“Embedment Properties of Headed Studs”, TRW Nelson Division, Lorain, OH.

I 3.30 Ollgaard, J. G., Slutter, R. G., and Fisher, J. W., “Shear Strength of Stud Con-
nectors in Lightweight and Normal Weight Concrete”, AISC Engineering Journal,
v. 8, no. 2, April, 1971.
I 3.31 McMackin, P. J., Slutter, R. G. and Fisher, 3. W., “Headed Steel Anchors under
Combined Loading”, AISC Engineering Journal, Second Quarter, 1973.

I 3.32 “Structural Welding Code, Dl.l-79”, American Welding Society, Miami, FL, 1979.

I 3.85
I
3.33 “Structural Welding Code/Reinforcing Steel”, D1.4-79”, American Welding Soci-
I
ety, Miami, FL, 1979.
3.34 “Standard Specifications for Highway Bridges‘I, Twelfth Edition, 1977, American I
Association of State Highway and Transportation Officials, Washington, DC.
3.35 “Manual of Steel Construction”, Seventh Edition, American Institute of Steel
Construction, Chicago, IL.
I
3.36 “Standard Specification for Deformed and Plain Billet-Steel Bars for Concrete
Reinforcement”, (ASTM A615-781, American Society for Testing and Materials,
I
Philadelphia, PA.
3.37 “Standard Specification for Low-Alloy Steel Deformed Bars for Concrete Rein-
forcement” (ASTM A706-761, American Society for Testing and Materials, Phila-
I
3.38
delphia, PA.
Conrad, R. F., “Tests of Grouted Anchor Bolts in Tension and Shear”, AC1 Journal,
I
v. 66, no. 9, Sept, 1969.
3.39 Anderson, A. R., ‘Composite Designs in Precast and Cast-in-Place Concretes, I
Progressive Architecture, v. 41, no. 9, Sept, 1960.
3.40 “Standard Specification for Uncoated Seven-Wire Stress-Relieved Strand for
Prestressed Concrete”, (ASTM A416-74), American Society for Testing and Mater-
I
ials, Philadelphia, PA.
3.41 “Standard Specification for Uncoated High-Strength Steel Bars for Prestressing
I
Concrete”, (ASTM A722-75), American Society for Testing and Materials, Phila-
delphia, PA.
I
3.42 “Standard Specification for Epoxy-Resin-Base Bonding Systems for Concrete”,
ANSI/ASTM C881-78, American Society for Testing and Materials, Philadelphia,
PA. I
3.43 Kriz, L.B. and Raths, C. H., “Connections in Precast Concrete Structures -
Strength of Corbels”, PC1 Journal, v. 10, no. l,Feb, 1965. I
3.44 Mattock, A. H., Chen, K. C. and Soongswang, K., “The Behavior of Reinforced

3.45
Concrete Corbels”, PC1 Journal, v. 21, no. 2, March-April, 1976.
Mattock, Alan H., “Design Proposals for Reinforced Concrete Corbels”, PC1
I
Journal, v. 21, no. 3, May-June, 1976.
3.46 Raths, Charles H., “Embedded Structural Steel Connections”, PC1 Journal, v. 19,
I
no. 3, May-June, 1974.
3.47 Marcakis, K. and Mitchell, D., “Precast Concrete Connections with Embedded
I
Steel Members”, PC1Journal, v. 25, no. 4, July-August, 1980.
3.48 Chan, Timothy, “A study of the Behavior of Reinforced Concrete Dapped-End
Beams”, Master of Science Thesis, University of Washington, 1979.
I
I
3.86
I
I
I 3.49 Cazaly, L., and Huggins, M., “Canadian Prestressed Concrete Institute Handbook”,
Canadian Prestressed Concrete Institute, 1964.
I 3.50 Ife, J., Uzumeri, S. and Huggins, M., “Behavior of the ‘Cazaly Hanger’ Subjected to
Vertical Loads”, PC1Journal, v. 13, no. 6, Dec., 1968.
I 3.51 Loov, Robert, “A Precast Beam Connection Designed for Shear and Axial Loads”,
PC1 Journal, v. 13, no. 3, June, 1968.

I 3.52 Gustaferro, A. II., and Martin, L. D., “PC1 Design for Fire Resistance of Precast,
Prestressed Concrete”, Prestreased Concrete Institute, Chicago, 1977.

I 3.53 “Corps of Engineers Specification for Non-shrink Grout”,


U. S. Army Corps of Engineers, 1978.
CRD-C588-78A,

3.54 AC1 Committee 503, IUse of Epoxy Compounds with Concrete”, AC1 Journal,
I v. 70, no. 9, Sept, 1973.
3.55 Mattock, A. II. and Gaafar, G. H., “‘Ihe Strength of Embedded Steel Sections as
I Brackets”, University of Washington, March, 1981. Presented at the 1981 Annual
Convention, American Concrete Institute, Dallas, TX, Feb, 1981.

I 3.56 Moustafa, S. E., “Effectiveness of Shear-Friction Reinforcement in Shear Dia-


phragm Capacity of Hollow-Core Slabs”, PC1Journal, v. 26, no. 1, Jan-Feb, 1981.

I
I
I
I
I
I
I
I
I
I 3.87
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I PART 4

I TYPICAL CONNECTION DETAILS

I
I
I
I
I
I
I
I
I
I
I
I 4. TYPICAL CONNECTION DETAILS

I The connection details shown in this section were selected from over 300 which
were initially reviewed. Some of these details are based on details shown in the
publications listed at the end of the section. Others were supplied by those who design
I and manufacture precast, prestressed concrete products.

Classification of Connections

I The connections have been classified by the types of members which are connected
as follows:

I 4.1
4.2
Column to foundation -
Column to column -
-
CF
cc
4.3 Beam to column BC
I 4.4
4.5
4.6
Slab to beam
Beam to girder
Beam to beam
-
-
-
SB
BG
BB
4.7 Slab to slab - ss
I 4.8
4.9
Wall to foundation
Slab to wall
-
-
WF
SW
4.10 Beam to wall - BW
I 4.11 Wall to wall - ww

Evaluation of Connections

I The more than 300 initial connection details were reviewed by approximately 40
professional, producer and associate members of the Prestressed Concrete Institute.
Each was initially evaluated on the basis of the criteria discussed in Section 3.1. This
I initial evaluation led to the selection of the details shown and the comments on each
detail.

I The final evaluation rating given for each connection is a composite, subjective
judgment based on the comments of the reviewers.

The first symbol in the “evaluation” rating refers to the extent of use in North
I America:

1- Common usage
I 2- Not common usage

The second symbol refers to the acceptability of each connection. Jn general, when
I applying this rating, it was assumed that the connection could be designed satisfactorily
to resist the loads imposed. Thus the evaluation was based more on simplicity, durability
and volume change accommodation than on inherent strength and ductility. The ratings
are as follows:
I Good solution
; - Acceptable
I AS
cs
-
-
Good solution for special design situations
Acceptable for special design situations
AM - Good solution when moment resistance is required

I CM - Acceptable when moment resistance is required.

4.1
I
I
I
Initially, there was also an “F” category for connections which, in the opinion of the
evaluators, had little merit. This accounts for the elimination of many of the initial
300. Others were combined with those shown, either by smaller inset drawings, or by
I
reference in the comments.

There was not universal agreement as to the merits of the details shown and there I
are probably many ‘good” and “acceptable” details not shown. As the state-of-the-art of
precast, prestressed construction continues to develop, other good connection ideas will
undoubtedly develop. I
I
I
I
I
I
I
I
I
I
I
I
I
I
4.2 I
I
I
I
I
I
I
I
I 4.1 Column to Foundation - CF

I
I
I
I
I
I
I
I
I
I
I 4.3
I
I
COLUMN TO FOUNDATION - CPl

The column is cast with four corner pockets and a base plate the same size as the
column. It is erected over anchor bolts protruding from the foundation. The space
I
between the column and foundation is filled with a dry-pack or non-shrink grout.

Temporary support and levelling are accomplished by tightening down on the nuts
I
with the column resting on a center stack of shims or by a doubIe nut (levelling nut)
arrangement as illustrated in the insets.

Features:
I
...Corner pockets allow easy wrench access and effective placement of
anchor bolts.
...Holes in base plate are oversized lo reduce tolerance problems.
. ..Column size base plate does not require form penetration and
I
permits the thinnest possible plates.
. ..Conne&ion is concealed and protected from corrosion after
patching.
I
. ..Bolting allows quick, easy erection in any weather.

Disadvantages: . ..Eliminates possibility


base plate.
of welding column corner reinforcing bars to I
Evaluation: . ..lA I
I
I
I
I
I
I
I
I
I
4.4 I
I
I COLUMN TO FOUNDATION - CF2

I This is a variation of connection CFI.


of temporary support.
Refer to CFl for description and discussion

I Features: . ..Side pockets allow corner bars to be welded to base plate, if desired.
. ..Oversized holes in base plate usually reduce tolerance problems.
. ..Column size base plate does not require form penetration and

I permits thinnest possible plate.


...Connection is concealed and protected from corrosion after
patching.
...Bolting allows quick, easy erection in any weather.
I Disadvantages: . ..Side pockets restrict wrench movement and provide less effective
placement of anchor bolts for moment resistance.
I Evaluation: . ..lC

I
I
I
I
I
I
I
I
I
I
I
I 4.5
I
COLUMN TO FOUNDATION - CP3
I
The base plate is larger than the size of the column and can be cast into the end as
shown or welded on as illustrated in the inset.
I
Refer to CFl for discussion and insets concerning erection and temporary support.
I
Features: ...No restrictions on wrench movement.
. ..Column corner reinforcing bars can be welded to plate for anchorage
if desired.
...Larger base plate increases effective bearing area if needed.
I
. ..Oversized holes usually reduce tolerance problems.
...Bolting allows quick, easy erection in any weather. I
Disadvantages: . ..Usually requires thicker base plate.
...Connection is not concealed or protected from corrosion.
. ..Large base plate may interfere with existing or future wall located I
near column line.
...Plate has to penetrate column form or be beyond end of form when
cast with column. Connection is not usually used with prestressed
columns cast in long line forms.
I
Evaluation: . ..lC I
I
I
I
I
I
I
I
I
I
4.6 I
I
I COLUMN TO FOUNDATION - CF4

I This is a variation of connection CF3. The base plate is wider than the column on
only two sides and can be attached in the same ways as CF3.

I Refer to CFl for discussion and insets concerning erection and temporary support.
Features: ...Savings on base plate material over CF3.
I . ..No restrictions on wrench movement.
...Main reinforcement can be welded to base plate if desired.
. ..Oversized holes usually eliminte tolerance problems.
. ..Larger base plate increases effective bearing area.
I . ..Bolting allows quick, easy erection in any weather.

Disadvantages: ...Requires somewhat thicker base plate.


I . ..Connection is not concealed or protected from corrosion.
. ..Plate has to penetrate column form or be beyond end of form when
cast with column. Connectioin is not usually used with

I Evaluation: . ..lC
prestressed columns cast in long line forms.

I
I
I
I
I
I
I
I
I
I
I
I
COLUMN TO FOUNDATION - CF5
I
This connection is similar in basic principle to CFl but uses two angles instead of a
base plate. Anchorage of angles in the column is provided by stiffeners and welded bars.
I
Refer to CF1 for discussion and insets concerning erection and temporary support. I
Features: ...Long pocket allows easy wrench access.
. ..Oversizad holes for bolts reduce tolerance problems.
...Angles are within dimensions of column which allows easy casting.
. ..Connection is concealed and protected from corrosion after
I
patching.
...Bolting allows quick, easy erection in any weather. I
Disadvantages: . ..Limitad by available angle thicknesses.
. ..More patching required.
. ..Eliminates possibility of welding main corner bar reinforcement to I
base angles.

Evaluation: . ..ZC I
I
I
I
I
I
I
I
I
I
I
4.8 I
I
I COLUMN TO FOUNDATION - CP6

I This is a variation of CF4 and is used when added base plate stiffness is needed.
The base plate is stiffened in the projecting direction. Anchorage is provided by
reinforcement welded to the stiffeners.
I Refer toCF1 for discussion and inset concerning erection and temporary support.

I Features: . ..Stiffeners allow a thinner base plate,


less rotation than CF4.
. ..Main reinforcement can be welded to
increased moment capacity and

stiffeners for anchorage if


desired.
I . ..No restrictions on wrench movement.
. ..Larger base plate increases effective bearing area.
...Bolting allows quick, easy erection in any weather.
I Disadvantages: . ..Connection is not concealed or protected from corrosion.
...Stiffeners and wide base plate make casting more difficult.

I Evaluation: . ..2cs

I
I
I
I
I
I
I
I
I
I
I 4.9
I
I
COLUMN TO FOUNDATION - CP7

This is a commonly used connection when moment resistance is required. Termed


“socket” connection, it involves sticking the column into a very rigid base and filling all
I
gaps with structural grout. This “locks in” the column and prevents rotation.

Shims and wedges are used as shown for temporary support and alignment.
I
Features: . ..Quick. easy erection in any weather.
. ..Moment resistance at column base.
. ..Minimum tolerance problems.
I
Disadvantages:
. ..Simplifies column casting.
. ..Foundation work is expensive.
I
...Difficult to assure good grouting under column.

Evaluation: . ..lAM I
I
I
I
I
I
I
I
I
I
I
I
4.10 I
I
I COLUMN TO FOUNDATION - Cl’8

I Termed “hammerhead” this connection requires a monolithically cast pedestal base


reinforced for the required moment capacity and containing sleeves at the anchor bolt
locations.
I This base is much stiffer
moment resistance.
than the average steel base plate and thus provides

I Features: ...Moment resistance at column base.


. ..Simplified foundation work.

I Disadvantages: ...Requires special form at column base, so is not suitable for long-line
casting.
...Possible tolerance problems with sleeves.
I Evaluation:
. ..Bolts not concealed or protected from corrosion.

. ..ZCM

I
I
I
I
I
I
I
I
I
I
I
I
I
I
COLUMN TO FOUNDATION - CP9

This is an acceptable, although not commonly used, connection for providing


moment resistance. The column has cast-in grout sleeves which fit over reinforcement
I
projecting out of the foundation. The sleeves are then grouted to affect a lap splice.
The gap under the column is filled with a dry-pack, non-shrink grout the same as previous
connections. The inset shows a proprietary sleeve with some advantages in production.
I
(See CC6).

Temporary lateral support must be provided by guying or other means of bracing I


until the sleeves are grouted and set.

Features: ...Moment resistance at column base


. ..Connection is concealed after patching.
I
...Often used for architectural columns where base is exposed.

Disadvantages: . ..Lap length for larger bars can be quite long.


I
...Requires additional means of temporary bracing.
. ..Requires accurate placement of projecting reinforcement.
...Reinforcement is easily bent before column is placed. I
Evaluation: . ..2CM
I
I
I
i
I
I
1
I
I
I
4.12 I
I
I COLUMNTOPOUNDATION-CCPlO

I The foundation has grout sleeves cast in which accept projecting reinforcement
from the column. The sleeves are filled just before the column is set. Dry-pack grout is
later worked under the column the same as the previous connections.
I The angles and bolts shown are one way of providing temporary support and
levelling until the grout sets. The angles can be reused or left in place.

I Features: ...Moment resistance at column base.


. ..Somewhat easier to fit bars into sleeves.

I Disadvantages: ...Anchorage length for larger bars can be quite long.


. ..Requires additional means of temporary bracing, such as the
hardware shown.
I ...Good weather is necessary when erecting column so sleeves can be
grouted.
. ..Precauticns must be taken to keep sleeves free of water and debris.

I Evaluation: . ..2CM

I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
4.2 Column to Column - CC
I
(Column splices)
I
I
I
I
I
I
I
I
I
I
4.14 I
I
I COLUMN TO COLUMN - CC1

I The column is cast with four corner pockets and a base plate usually somewhat
smaller than the column. It is erected over anchor bolts protruding from the column
below. The space between columns is filled with a dry-pack or non-shrink grout.

I Temporary support and levelling are accomplished by tightening down on the nuts
with the column resting on a center stack of shims or by a double nut (levelling nut).
(See CFl insets).
I Features: ...Corner pockets allow easy wrench access.
. ..Holes in the base plate are oversized to reduce tolerance problems.
I . ..Connection is concealed and protected from corrosion after
patching.
. ..Bolting allows quick, easy erection in any weather.

I Disadvantages: . ..Eliminates possibility of welding main corner bar reinforcement to


base plate.
. ..Anchor bolts have limited moment capacity, so connection should be
I placed near inflection point.

Evaluation: . ..lA
I
I
I
I
I
I
I
I
I
I
I 4.15
I
COLUMN TO COLUMN - CC2
I
This is a variation
of temporary support.
of connection Ccl. Refer to CC1 for description and discussion I
Features: ...Side pockets allow corner bars to be welded to base plate.
. ..Connection is concealed and protected from corrosion after
patching.
I
Disadvantages:
...Bolting allows quick, easy erection in any weather.

. ..Side pockets restrict wrench movement and provide less effective


I
placement of anchor bolts for moment resistance.
. ..Anchor bolts have limited moment capacity, so connection should be
placed near inflection point.
I
Evaluation: . ..lC
I
I
I
I
I
c
I
5
I
I
I
I
I
I
4.16 I
I
I COLUMN TO COLUMN - CC3

I This connection is similar in basic principle to CC1 but uses two angles instead of a
base plate. Anchorage of angles in the columns is provided by stiffeners and welded bars.

I Refer to CC1 for erection and temporary

Features:
support.

...Long pocket allows easy wrench access.

I . ..Oversized holes for bolts reduce tolerance problems.


...Connection is concealed and protected from corrosion after
patching.
...Bolting allows quick, easy erection in any weather.
I Disadvantages: . ..Axial capacity and moment capacity are limited by available angle
thicknesses.
I . ..More patching required.
. ..Anchor bolts have limited moment capacity.

I Evaluation: . ..2c

I
I
I
I
I
I
I
I
I
I
I 4.17
I
I
COLUMN TO COLUMN - CC4

This connection is commonly used when a moment transfer splice is desired. The
columns are match cast with top and bottom plates and then welded together when
I
erected.

Features: ...Moment resistance at connection.


I
-Field fitting problems are minimized if correctly match cast.
. ..Immediate full bearing so erection can proceed to upper levels
without delays for grouting.
...Connection is concealed and protected from corrosion after
I
Disadvantages:
patching.

. ..Match casting requires special care in the plant.


I
. ..A significant amount of welding is required.
. ..No “quick connection” for erection at column base as in bolted
connections.
I
. ..Correction of minor errors is difficult.

Evaluation: . ..lCM I
I
I
I
I
I
I
I
I
I
I
4.18 I
I
I COLUMN TO COLUMN - CC5

I The lower column has grout sleeves cast in which accept projecting reinforcement
from the upper column. The sleeves are filled as the column is erected. When set, it
affects a lap splice. Dry-pack grout is later worked under the column.
I Temporary support and levelling must be accomplished by guying or other means of
bracing until the sleeves are grouted and cured.

I Features: ...Moment resistance at connection.


. ..Usually few tolerance problems.
. ..Somewhat easier to fit bars into sleeves than CC7.
I . ..Connection is protected from corrosion and usually acceptable for
exposed columns.

I Disadvantages: . ..Lap length for larger bars can be quite long.


. ..Requires additional means of temporary bracing.
. ..Good weather is necessary when erecting columns for sleeves to be
grouted.
I ...Precautions must be taken to keep sleeves free of water and debris.
. ..Main reinforcement in lower column must be bent to avoid sleeves.

I Evaluation: . ..lCM

I
I
I
I
I
I
I
I
I
I
I
COLUMN TO COLUMN - CC6
I
“ihii is a proprietary mechanical bar splicing system that is gaining popularity
world-wide. Special installation devices are provided by the manufacturer to aid install-
ation.
I
In the preferred installation, bars project from the lower column and the sleeves fit
over the top. Leveling is accomplished by shims. Grout is pumped into the bottom tube
I
until it comes out the top. In the alternate shown in the inset, the bars project from the
upper column section. Grout may be placed in the sleeves and the shim space in the
same operation, prior to placing the top section. I
Features: . ..Moment resistance at connection.
. ..Special installation devices reduce tolerance problems.
. ..No patching required.
I
. ..Column bars need not be bent.
. ..Effectiveness of proprietary sleeves have been verified by test. I
Disadvantages: . ..Additional means of bracing must be provided until grout in sleeves
has set.
. ..Requires a proprietary mechanical sleeve. I
. ..2A
Evaluation:
I
I
I
I
I
I
I
I
I
I
4.20
I
I
I COLUMN TO COLUMN - CC7

I The upper column has grout sleeves cast in which fit over reinforcement protruding
from the lower column. The sleeves are then grouted to affect a lap splice. The space
between column sections is filled with a dry-pack grout.
I Temporary support and levelling must be accomplished by guying or other means of
bracing until the grout in the sleeves is cured.

I Features: ...Moment resistance at connection.


. ..Connection is concealed after patching.

I Disadvantages: ...Lap length for larger bars can be quite long.


. ..Requires additional means of temporary bracing.
. ..Placement of reinforcement and tubes is critical.
I . ..Main reinforcement in upper section must be bent around the
sleeves.
. ..Protruding reinforcement can be bent during handling.

I Evaluation: . ..2CM

I
I
I
I
I
I
I
I
I
I
I 4.21
I
COLUMN TO COLUMN - CC8
I
This connection is only used when full moment transfer is desired in large, heavily
reinforced columns. Both column sections are cast with reinforcement protruding from
I
the ends which is welded together when erected.

Temporary support and levelling are accomplished by resting the column on a I


center stack of shims and welding several of the bars together. The entire connection is
later grouted.

Features: . ..Moment resistance at connection.


I
. ..Long lap lengths on reinforcement is not required.
. ..Connection is concealed and protected from corrosion after
patching.
I
Disadvantages: . ..Reinforcement
. ..Reinforcement
must be precisely placed to accomplish welding.
must be a weldable grade steel. I
Evaluation: . ..ZCM
I
I
I
I
I
I
I
I
I
I
I
4.22 I
I
I COLUMN TO COLUMN - CC9

I This connection may be used when fully continuous, composite ductile frames are
required. It allows the placement of column ties for confinement. Beam reinforcement
is placed through the joint and the assembly is completed with cast-in-place concrete.
I Temporary support is provided by a sleeve-in-sleeve connection. A structural pipe
or tube protruding from the upper section fits into a slightly larger pipe or tube in the
I lower section. A small weld holds the assembly in place. No weld is required if the fit is
tight enough.

Features: ...Moment resistance at connection (beam and column).


I . ..Next level can be erected before connection is poured and cured.
. ..A ductile connection.
. ..Connection is concealed and protected from corrosion.
I Disadvantages: ...Sleeves must be very accurately placed in the plant.
. ..Connection is congested, so extreme care is necessary to avoid

I honeycombing in cast-in-place connection.


. ..Requires much field labor.

Evaluation: . ..2CM
I
I
I
I
I
I
I
I
I
I
I 4.23
I
COLUMN M COLUMN - CC10
I
‘This system uses vertical post-tensioning for the column reinforcement and for
splicing the column sections. Sleeves are cast in the columns at the plant. The tendons I
knmally bars) are attached to an anchor (at the bottom floor) or a coupler (at intermed-
iate floors). The upper column is then “threaded” over the bars. The bars are tensioned
and anchored, leaving enough projection to attach a coupler to receive the bars for the
next level.
I
Features: . ..Ductile moment resisting connection.
. ..Post-tensioning reduces drift in high-rise buildings.
I
Disadvantages: . ..Complex erection procedure.
. ..Post-tensioning is an added operation.
. ..Alignment of sleeves is critical.
I
...Requires supplemental reinforcement or pretensioning for handling.
Evaluation: . ..2CM
I
I
I
I
I
I
I
I
I
I
I
I
4.24
I
4.3 Beam to Column - BC

4.25
I
BEAM TO COLUMN - BCl
I
The beam sits on a bearing pad (usually neoprene or laminated fabric) which pro-
vides even bearing and permits slight movements caused by shrinkage, creep and temper-
I
ature change. The top connection transfers horizontal shear forces between the beam
and column, provides erection stability and braces the column, but does not provide
rotational restraint. I
The top connection should allow some slight rotation to avoid attracting unintended
negative moments. Thus
welded. The inset on the
when the angle connection is used, only the outer ends are
left shows an alternate top connection which permits rotation
I
about the horizontal bar. The inset on the right illustrates a dapped-end alternate.

Refer to Sect. 3.5.1 for column corbel design and 3.5.3 for dapped-end beam design. I
Features: . ..Quick. easy erection.
. ..Few tolerance problems.
. ..Minimum volume-change restraint.
I
Disadvantages: ...No moment capacity.
. ..No torsional restraint.
I
Evaluation: . ..lA
I
I
I
I
I
I
I
I
I
I
4.26
I
I
I BMM WI COLUMN - BC2

I A sleeve in the end of the beam fits over a dowel which is threaded into either an
insert in the support or a ferrule welded to the underside of a bearing plate. To prevent
damage in handling, the dowel is put in just prior to erection. The sleeve should be 3 or 4
I times the size of the dowel to minimize field tolerance problems.
In order to prevent restraint, the bottom few inches of the sleeve should be filled
with a compressible material such as sand, vermiculite, or asphalt. The remainder is
I filled with grout.
The inset shows an alternate threaded dowel with a plate washer and nut when
I torsion or uplift resistance is desired. It also illustrates use with a dapped-end beam.
Features: . ..Quick. easy erection if properly detailed.
. ..Volume change restraint is minimized.
I . ..Provides shear resistance and some torsional restraint after
grouting.

I Disadvantages: . ..No lateral connection until sleeves are grouted.


. ..No reliable moment capacity.
. ..Sleeve can fill with water during erection and then freeze and crack
I Evaluation: ...lA
the beam, unless precautions are taken.

I
I
I
I
I
I
I
I
I
I
I
BEAM TO COLUMN - BC3
I
This is a variation of connection BC2. The dowels are shown as cast in and
protruding from the top of the column. Inserts or ferrules, as described in BC2, can also
I
be used.

The inset shows two dowels with threaded ends for each beam when increased
I
torsional resistance is required. The nuts and washers can be recessed and patched if
desired.

Separate bearing pads under each beam are required and beam bearing plates may
I
be necessary.

Features: ...Quick. easy erection.


I
. ..Volume change restraint is minimized.
...Few tolerance problems if sleeve is large enough.
I
Disadvantages: . ..No lateral connection until sleeves are grouted.
. ..No moment capacity.
. ..Sleeve can fill with water and cause damage if frozen. I
Evaluation: . ..lC
I
I
I
I
I
I
I
I
I
I
4.28 I
I
I BEAM TO COLUMN - BC4

I A wide flange, tube or other structural steel section is embedded in the top of the
column and fits in vertical slots in the ends of the beams. Plates are welded to the top
of the steel section and to plates or angles in each beam that can be recessed and
I patched as shown. See BCl for comments on design of top connection.

Separate bearing pads under each beam are required and beam bearing plates may

I be necessary.

Features: . ..Immediate erection connection.


. ..No tolerance problems.
I . ..Torsional resistance provided during erection and for service loading.

Disadvantages: . ..Protruding steel section causes difficulties in casting.


I Evaluation: . ..2c

I
I
I
I
I
I
I
I
I
I
I
I 4.29
I
BEAM TO COLUMN - BC5
I
BC5 through BClO show several variations of connections which use structural steel
members projecting from the column to support the beams. Design of these connections
is discussed in Sect. 3.5.2.
I
In this version, the exposed steel haunch usually requires encasement in concrete,
or other protection (See Sect. 3.1.4 - 3.1.5).
I
The h-metshows attachment of the projecting member by welding to avoid penetra-
tion of molds. This is discussed in Sect. 3.1.6. I
See BCl for top connection
Features: . ..High load carrying capacity.
I
. ..Reduced structural depth compared with concrete corbel.
. ..Minimum forming.
. ..Volume change restraint minimized.
I
Disadvantages: . ..Requires encasement for fire and corrosion protection.
. ..Alignment during casting is critical. I
Evaluation: . ..lA

c
I
I BEAM TO COLUMN - BC6

I (See BC5). In this version the haunch is recessed to reduce the depth and the
amount of patching required for fire and corrosion protection.

I Features:

Disadvantages:
. ..Clean looking, concealed connection (See BC5).

. ..Requires overhead patching to complete encasement.


. ..Beam end requires dapped-end design (See Sect. 3.5.3)cSee BC5).
I Evaluation: . ..lA

I
I
I
I
I
I
I
I
I
I
I
I
I
I 4.31
I
BEAM TO COLUMN - BC?
I
This connection employs two angles which form a “cradle” making the beam
immediately stable when erected. Bearing pads and a top connection as discussed in BCl
I
are required. The angles can be recessed and patched if required.

Features: . ..Quick. easy erection.


. ..Clean looking, concealed connection.
I
. ..Some adjustment for tolerances.
. ..Few volume change restraint problems.
...Provides stability during erection.
I
Disadvantages: . ..Alignment of angles during casting is critical.
...Angles have limited capacity-connection cannot be used for heavy
I
loads.

Evaluation:
. ..Requires form penetration.

. ..2A
I
I
I
I
I
I
I
I
I
I
I
I
4.32
I
I
I BEAM TO COLUMN - BC8

I This is another variation of the embedded steel haunch discussed in BC5. It is a


built-up section out of steel plates which increases versatility. The haunch can be made
wide enough to avoid the rolling at erection and increased bearing that could be a
I problem with some of the other steel haunches shown. Patching to conceal the
connection is possible, but may be more difficult.

I Features:
Disadvantages:
...(See BC5).
. ..Built-up steel section requires additional labor.
. ..Patching to conceal connection may be difficult.
I . ..Requires dapped-end beam design.
Evaluation: . ..2A
I
I
I
I
I
I
I
I
I
I
I
I
I 4.33
I
BEAM TO COLUMN - BC9
I
See BC5. This connection employs two steel channels separated for a bolt to go
through. A sleeve is cast in the beam and recessed pockets allow the connection to be I
completely concealed after patching.

The bolt and sleeve offer a weldless top connection which provides lateral stability
when grouted. A bearing pad is also required.
I
Features: ...Quick. easy erection.
. ..No welding required.
I
. ..Concealed connection.

Disadvantages: . ..(See BCS). I


. ..Requires dapped-end beam design
Evaluation: . ..ZA I
I
I
I
I
I
I
I
I
I
I
I
4.34
I
I
I BEAM TO COLUMN - BClO

I (See BC5). An angle is welded to a plate cast in the column. For light loads, the
plate can be attached to the column with headed studs. Stiffeners on the angle, as
illustrated in the inset, can increase the capacity, but for the most part, the connection
I is limited to light loads.

Features: . ..Quick. easy erection.

I . ..No column form penetrations.


. ..Provides stability during erection, especially with stiffeners
shown in inset.
as

I Disadvantages: . ..Limited to light loads.


. ..Considerable welding required.

I Evaluation: . ..zc

I
I
I
I
I
I
I
I
I
I
I
I
I
BEAM TO COLUMN - BCll
I
This connection is especially suitable for repair or remodeling. It is not commonly
used for conventional precast construction. Holes are drilled or sleeves are cast in the
I
column through which bolts pass. An epoxy grout is injected to completely fill the
sleeves or drilled holes. Oversized or plate washers are used at each end.

A bearing pad and top connection as discussed in BCl are required.


I
Features: ...Can be used for repair or corrective measures.
. ..Lips formed by stiffeners can act as “keepers” when first erected.
I
Disadvantages: ...Limited to light loads.
. ..Steel alternate requires encasement for fire and corrosion I
protection. -
Evaluation: . ..2AS I
I
I
I
I
I
I
I
I
I
I
I
4.36
I
I
I BBAM To COLUMN - BC12

I BC12 through BC14 show several variations of hanger connections. Design of these
connections is discussed in Sect. 3.5.4. Hanger connections are immediately stable
during erection, although without a bottom stopper, may “roll” when loaded eccentri-
I cally. Main drawing shows beam framing along the centerline of columns and inset shows
beam framing into the side of the columns.
Other variations of steel haunches are shown in BC5 through BClO and design is
I discussed in Sect. 3.5.2.
Features: ...Quick. easy erection.
I . ..Erection stability.
. ..Concealed connection
Disadvantages: . ..Connection hardware is expensive.
I . ..Welding of reinforcement is critical.
. ..Alignment of the embedded haunch is critical.

I Evaluation: . ..2c

I
I
I
I
I
I
I
I
I
I
I
I
BEAM TO COLUMN - BC13
I
This connection is similar to BC 12, except the long tube haunch fitting in the beam
slot provides torsional stability. It is especially suitable for eccentrically loaded I,- I
shaped beams.

Features: -(See BC12). I


. ..Torsional stability.
Disadvantages: . ..(See BC12).
Evaluation: . ..ZAS

4.38
I
I BEAM TO COLUMN - BC14

I This is an acceptable beam to column connection for special situations such as for
shallow beams where headroom is a problem. It is not commonly used for conventional
precast construction.
I Features: . ..Quick. easy erection.
. ..Good field adjustment.
. ..No volume change restraint problems.
I . ..Connection is concealed and protected
is poured.
from corrosion after topping

I Disadvantages: ...Column is not braced until topping is poured.


. ..May have a tendency to roll until topping is poured.

I Evaluation: . ..2cs

I
I
I
I
I
I
I
I
I
I
I
4.39
I
I
I
BEAM TO COLUMN - BC15

This is a commonly used connection when moment resistance is desired. Plates in


the top and bottom of the beam are welded to plates in the columns. The connection
I
must be designed to resist all forces including those from volume change restaint.

Features: . ..Quick. easy erection.


I
. ..Moment resistance at connection.

Disadvantages: ...Connection is exposed and may need protection.


. ..Volume change restraint must be considered.
I
Evaluation: . ..lCM I
I
I
I
I
I
I
I
I
I
I
I
I
4.40
I
I
I
BEAM TO COLUMN - BC16

I BC16 through BC19 show several variations of composite moment connections.


These are the most commonly used connection types in ductile frames as discussed in
sect. 2.
I In this version, the beams bear on a “hammerhead” column. Bearing plates in the
beam are welded to angles in the column haunches, which are connected together with

I welded reinforcement. The tension steel is placed and the composite topping is cast.
After the topping has cured, the next column is erected over the projecting reinforcing
bars.

I The inset shows an alternate


reinforcement.
method of providing continuity in the bottom

I Features: . ..Full moment resistance at the connection.


. ..Easy field adjustment.

I Disadvantages: ...Slow erection, as composite pour must cure before next column is
placed.
. ..Overhead welding is difficult.

I . ..Exposed steel may require fire or corrosion protection (See


sect. 3.1.4 - 3.1.5).
. ..(Inset) Requires reinforcement projecting from beam end.

I Evaluation: . ..lAM

I
I
I
I
I
Ii
I
I 7
I
I
I
I
BEAM TO COLUMN - BC17

This is similar to BC16 except the column corbels are omitted. Temporary shoring
of the precast beam is required. Bottom reinforcement continuity is attained by lapping,
I
welding, or hooking (See BC16 inset) projecting reinforcement, depending on dimensions
available. I
Features: . ..Clean connection with no projecting corbels.
. ..No exposed steel.
. ..(See also BC16).
I
Disadvantages: ...Requires temporary
. ..(See also BC16).
shoring during erection.
I
Evaluation: . ..lAM
I
I
I
I
I
I
I
I
I
I
I
I
4.42
I
I
I BEAM TO COLUMN - BC18

I The column is a multi-story precast column with reinforced concrete corbels. It is


shown here as rectangular and being used with a soffit beam system. (See inset). The
beams are set and the base plates are welded to angles in the haunch. (Alternates as
I shown on BC16 and BC17 can also be used). The negative moment steel is placed through
sleeves in the column. The composite cast-in-place topping is poured after deck
members are erected.

I Features: ...Full moment resistance at connection.


. ..Multi-story columns are more economical.
. ..Next floor can be erected before cast-in-place concrete work is
I completed.

Disadvantages: . ..Bottom connection (angles and plates) may require fireproofing.


I . ..Overhead weld is difficult.
. ..Possible tolerance problem with sleeves.
. ..Shallow soffit beams usually require temporary shoring.

I Evaluation: . ..lAM

I
I
I
I
I
I
I
I
I
I
I 4.43
I
BEAM TO COLUMN - BC19
I
This is a moment connection using precast beams and cast-in-place columns. The
beams have reinforcement projecting from the ends, which is lapped with horizontal
reinforcement in the connection. The beams require shoring. Each pour typically ends at
I
the top of the finished floor.

Features: ...Full moment resistance at connection.


I
. ..Easy adjustment for tolerances.
. ..Utilizes inexpensive precast beams and avoids expensive precast
columns.
I
. ..Connection is concealed.

Disadvantages:
. ..Lapped splices on reinforcement easier than welding.

. ..Precast erection must wait for cast-in-place pour.


I
Evaluation:
. ..Shoring is required.

. ..ZCM
I
I
I
I
I
I
I
I
I
I
I
I
4.44
I
I
I BEAM TO COLUMN - BC20

I A post-tensioning tendon is fed through a duct in the beam and an oversized sleeve
in the column. An anchorage plate is attached at the anchorage pocket and the tendon is
tensioned from the other end and locked off in the recessed pocket provided. Prior to
I tensioning, the space between the beam and column is filled with a dry-pack grout.

The inset illustrates the connection with a beam from both sides. The column can

I be haunched or pocketed as desired.

Features: ...Moment resistance at connection (negative moment only).


. ..Connection is concealed.
I Disadvantages: ...Anchorage bearing stresses must be considered in the column design.
. ..No positive connection until tendons are jacked unless other means
I Evaluation:
are provided.

. ..lAM

I
I
I
I
I
I
I
I
I
I
I
4.45
I
I
BEAM TO COLUMN - BC21 I
This shows a connection at the top of a column. Unlike BC20, the main
reinforcement for the beam and column is the post-tensioning.
As with BC20, the post-tensioning tendons are fed through embedded ducts
I
(oversized sleeves when passing through a second member) and are tensioned after the
grout between members has set. I
This connection can be designed to provide full moment resistance for lateral
loads. For earthquakes, the tendons can be positioned to establish hinging at a pre-
determined location (See Sect. 2). I
Features: ...Moment resistance at connection.
. ..Connection is concealed. I
Disadvantages: . ..No positive connection until tendons are jacked unless additional
means are provided.
. ..Without supplemental reinforcement or pretensioning, the beams
I
must be shored.

Evaluation: . ..lAM I
I
I
I
I
I
I
I
I
I
4.46
I
I
I
I BEAM TO COLUMN - BC22

I This connection is usually used with L-beams in parking structures. The beams set
on bearing pads in a pocket in the column, or on a corbel. They are bolted through
sleeves in the column into inserts in the beams. The bolts are located near the top or

I bottom of the beams depending on the direction of torsional rotation from eccentric
loads. The other (top or bottom) usually has a compression pad or shim between the
beam and column.

I Features: ...Torsional resistance provided.


. ..Connection is concealed and protected from fire or corrosion after
patching.
I Disadvantages:
...Few volume change restraint problems.

. ..Location and alignment of sleeves and inserts is critical.

I Evaluation: ...lAS

I
I
I
I
I
I
I
I
I
I
I
4.47
I
I
BEAM TO COLUMN - BC23
I
This is a commonly used connection for spandrel beams. The beams sit on an
embedded steel haunch and are bolted into the column, top and bottom. The clip angles
I
have a vertical slot ins one leg and a horizontal slot in the other for adjustment,
A bearing pad as discussed in BCl is required.
and patched.
The angles may be set in recesses I
Features: . ..Torsional resistance provided.
. ..No welding required.
I
Disadvantages: ...Connections are exposed and may require fire or corrosion
protection. I
. ..Alignment of inserts is critical.
. ..To avoid volume change restraint, there must be some slot remaining
after adjustment and bolts should not be overtightened. I
Evaluation: . ..lCS
I
I
I
I
I
I
I
I
I
I
I
4.48
I
I
I BEAM TO COLUMN - BC24

I This is an acceptable beam to column connection for special situations such as


cantilevers. The beam is erected over post-tensioning rods which have been placed in
conduits in the column and coupled to the rods from the lower level at the pocket
I provided. The beam is set on shims and then grouted underneath. When the grout has
set, the rods are tensioned and locked off. Then, the next column is erected and grouted.
Features: ...Full moment connection.
I . ..Cantilevered beam passes through uninterrupted.
...Connection is concealed and protected from corrosion after
patching.
I Disadvantages: ...No immediate connection at erection.
. ..Erection must wait for grouting and tensioning.

I Evaluation:
. ..Possible alignment problems with conduits.

. ..2cs

I
I
I
I
I
I
I
I
I
I
I
4.49
I
I
I
BEAM TO COLUMN - BC25

This is an acceptable
cantilevers. The columns
beam to column connection for special situations such as
have pockets and plates similar to connection Ccl. After the
I
beam is erected on shims,
Enough rod is left sticking
as described in connection
the threaded rods are inserted and the sleeves are grouted.
up so that the next column can be erected in the usual manner
Ccl.
I
Features: . ..Quick. easy erection.
. ..Cantilevered beam passes through uninterrupted.
I
...Connection is concealed and protected from corrosion after
patching.
...Usually no tolerance problems. I
. ..Alignment of sleeves is critical.
Disadvantages:
. ..Projecting bolts are easily damaged. I
Evaluation: . ..lCS
I
I
I
I
I
I
I
I
I
I
I
4.50
I
I
I
I
I
I
I
I
I 4.4 Slab to Beam - SB

I
I
I
I
I
I
I
I
I
I
I 4.51
I
SLAB TO BEAM - SBl
I
Standard connection when no mechanical tie to the support is required. Slabs bear
on high density plastic or hardboard bearing strips. Composite topping often used with
I
this connection.

Features: ...Quick. easy erection.


I
. ..Aflows adequate tolerance.

Disadvantages:
. ..Volume change restraint minimized.

. ..Without topping, no positive tie to beam for shear transfer or torsion


I
Evaluation: . ..lA
restraint.
I
I
I
I
I
I
I
I
I
I
I
I
I
4.52
I
I
I SLAB TO BEAM - SB2

I This connection provides a positive integral floor system without welding or cast-
in-place topping. With projecting stirrups as shown, the cast-in-place top portion of the
beam acts compositely with the precast section, providing added structural capacity.
I Features: ...Positive connection to beam.
. ..No hardware cast into slabs.
I Disadvantages:
. ..Allows adequate tolerance.

. ..Placement of longitudinal bars along top of beam requires threading


through the stirrups.
I . ..No positive tie until joints are grouted and field concrete is cast.
...Weather sensitive.

I Evaluation: . ..lA

I
I
I
I
I
I
I
I
I
I
I
4.53
I
I
SLAB TO BEAM - SB3
I
This is a standard connection for double tee roofs. The stems set on bearing pads
td a top connection is made as shown. Although one top connection per tee will usually
I
rffice for erection and diaphragm forces, two may be desirable. Experience has shown
tat in most stemmed members, a welded top connection will not cause volume change
?staint problems if the bottom of the stem is not restrained.
I
Features: ...Quick. easy erection.
. ..Allows adequate tolerance.
...Usually. no volume change restraint problems.
I
. ..Positive connection for erection and lateral forces.
Evaluation: . ..lA
I
I
I
I
I
I
I
I
I
I
I
I
I
4.54
I
I
I SLAB TO BEAM - SB4

I Precast slabs sit on a high density plastic or hardboard bearing strip. A loose plate
is welded to embedded plates in the slab and beam. A roof spandrel condition is
illustrated here.
I Features: ...Quick. easy erection from top.
. ..Positive connection to beam for erection and lateral forces.

I Disadvantages:
. ..Allows adequate tolerance.

. ..Hardware embedment difficult for extruded type voided slabs.

I Evaluation: . ..lA

I
I
I
I
I
I
I
I
I
I
I
I
4.55
I
I
SLABTOBEAM-SB5
I
This connection is used to transfer lateral diaphragm loads and to provide bracing
for the compression flange of steel beams. The inset shows a connection used when
I
maximum clearance is needed below the slabs. (Top flange bracing not provided in this
case.)

Features: ...Allows adequate tolerance.


I
. ..Provides bracing for beam flange.

Disadvantages: -Embedded hardware difficult with extruded type hollow-core slabs.


I
. ..Overhead welding is difficult.

Evaluation: . ..lC I
I
I
I
I
I
I
I
I
I
I
I
I
4.56
I
I
I SLABTOBEAM-SB6
I As with SBI,
hardboard bearing
precast slabs, generally voided, sit on a high density plastic or
strip. Steel plates with deformed bar anchors lay in the grout joints
and are welded to angles in the beam. The connection is completed as the joints are
I normally grouted. Usually, no topping used with this detail.

Features: . ..Positive connection to beam.


I ...No hardware cast into slabs.
. ..No cast-in-place concrete required.

Disadvantages: ...No connection until slabs are grouted.


I . ..Requires accurate, pre-planned placement of beam hardware.

Evaluation: . ..2c
I
I
I
I
I
I
I
I
I
I
I
I
I 4.57
I
SLAB TO BEAM - SB7
I
This connection used when a positive tie across the beam is desired, but there is no
grout key for embedment of reinforcing bars. Continuity for live loads (and dead loads if
I
slabs are shored before welding and grouting) can be provided if end joint is grouted.

Features: ...Quick. easy erection. I


. ..Allows adequate tolerance.

Disadvantages:
. ..Can provide continuity.

. ..Requires embedded hardware and blockout in slab.


I
...Slabs must be reinforced for intended or unintended negative
moments. I
Evaluation: . ..lCS
I
I
I
I
I
I
I
I
I
I
I
I
4.58
I
I
I SLAB TO BEAM - SB8

I This shows one variation of a hanger connection. It is used when structural


must be kept to a minimum. Design of hangers in discussed in Sect. 3.5.4.
depth

I Features: . ..Quick. easy erection.


. ..No pockets or ledge in beam.
. ..Allows adequate tolerance.

I Disadvantages: ...Usually limited to light loads.


. ..Hardware may be expensive.

I Evaluation: . ..ZCS

I
I
I
I
I
I
I
I
I
I
I
I
4.59
I
I
I
I
I
I
I
I
4.5 Beam to Girder - BG I
I
I
I
I
I
I
I
I
I
I
4.60
I
I
I BEAM M GIRDER - BGl

I The beam or joist sits on a bearing pad in a pocket in the girder. Depending on the
relative depths of the beam and girder, the~beam may need to be dapped to allow
adequate depth under the pocket.
I The beams and girders are often used in conjunction with a floor system having a
composite topping. If there is no topping, a top connection may be required.

I Features: ...Quick. easy erection.


. ..No dust-collecting ledges between joists.
. ..Minimum volume change restraint.
I . ..Beam has direct bearing.

Disadvantages: ...Difficult to correct if beams and pockets don’t align.


I Evaluation:
. ..Effective section of girder is reduced.

. ..2c

I
I
I
I
I
I
I
I
I
I
I
4.61
I
I
BEAM To GIRDER - BG2
I
The hanger connection shown is similar in concept to BC12 - BC14. Design of
hangers is discussed in Sect. 3.5.4.
I
Features: . ..Quick. easy erection.
...Minimum reduction of girder section. I
. ..No dust-collecting ledges.

Disadvantages: ...Alignment of pockets is critical.


. ..Beam hardware expensive.
I
Evaluation:
. ..Welding of reinforcement is critical.

. ..2A
I
I
I
I
I
I
I
I
I
I
I
I
I
4.62
I
I
I BEAM TO GIRDER - BG3

I A doweled connection with features similar to BC2.

Features: ...Quick erection.

I . ..Positive connection.
. ..Full section of girder is effective.

Disadvantages: . ..Structural depth of system.


I ...Dust can collect on top of girders.
. ..Alignment of dowels critical.
...Sleeves can fill with water and possibly freeze.
I Evaluation: . ..2c

I
I
I
I
I
I
I
I
I
I
I
I
4.63
I
I
BRAM TO GIRDER - BG4
I
This drawing shows a composite system using precast beams or joists on a soffit
beam. As shown, removable forming is required between beams, but a similar system
I
using hollow-core, flat slab or stemmed deck members is also used.
Features: ...Quick. easy erection. I
. ..Shallow structural depth.

Disadvantages:
. ..Totally integrated structural framing system.

. ..Composite system is weather dependent.


I
...Soffit of girder requires temporary shoring.

Evaluation: . ..lA
I
I
I
I
I
I
I
I
I
I
I
I
I
4.64
I
4.6 Beam to Beam - BB

4.65
I
BEAMTOBEAM-BBl
I
BBl through BB3 are used in continuous frames, where it is desired to have the
connection away from the column. Examples are “tree” columns with “drop-in” beams,
I
cruciform beam-columns with connections at mid-span, or systems which use story-high
columns and continuous beams.

This connection requires dapped-end designs on both segments (See Sect. 3.5.3).
I
Modification of hanger connections shown in BC12 - BC14 can also be used. The top
connection should be detailed as discussed in BCl. I
Features: . ..Quick. easy erection.

Disadvantages: . ..Dapped ends require congested reinforcement.


I
. ..No moment capacity.

Evaluation: . ..2c I
I
I
I
I
I
I
I
I
I
I
I
4.66
I
I
I BEAM TO BEAM - BB2

I ‘IT& connection is usually used in composite systems when “drop-in” beam sections
are suspended from two cantilevers. The connector is attached to the drop-in section.
Features: . ..Quick. easy erection.
I . ..Concealed connection.
Disadvantages: . ..Limited capacity.
I Evaluation: . ..X

I
I
I
I
I
I
I
I
I
I
I
I
I 4.67
I
I
BEAM TO BEAM - BB3
I
This connection is similar to BC6 with two embedded angles forming a “cradle”. An
alternate with a weld-on haunch is shown in the inset. Bearing pad(s) and top connections
I
are required and the haunches can be recessed and patched if desired.

Features: ...Quick. easy erection. I


Disadvantages: . ..Connection not as easy to conceal and protect as BBl or BBZ.
. ..No moment capacity. I
Evaluation: . ..2c
I
I
I
I
I
I
I
I
I
I
I
I
I
4.68
I
4.7 Slab to Slab - SS

4.69
I
sLABTosLAB-SSI I
This is the standard connection used between hollow-core and sohd slabs. The size
and shape of the key vary with the method of manufacture. The key is usually filled with
a sand-cement grout.
I
This connection distributes vertical loads and provides horizontal shear transfer for
moderate loads when the deck is used as a diaphragm. This is discussed in Sect. 2.
I
Features: ...Quick. easy erection.
...No embedded items in slabs.
I
Disadvantages:
. ..No hardware to corrode.
. ..Susceptible to freezing.
I
. ..Underside may have to be caulked if exposed.
Evaluation: . ..lA I
I
I
I
I
I
I
I
I
I
I
I
4.70
I
I
I SLAB TO SLAB - SS2

I An angle with deformed bar anchors and studs is embedded in the slab. A loose
plate is welded to the angles.

I The connection can be recessed slightly if the slab is thick enough to accommodate
the vertical leg of the angle or the angle can be inverted as shown in the inset.

I Features: ...Quick. easy erection.


.-Properly spaced connections distribute
diaphragm forces.
loading and transfer
. ..Helps even out differential camber.
I Disadvantages: ...More material per connection than SS5.

I Evaluation: . ..lA

I
I
I
I
I
I
I
I
I
I
I
I
I
SLAB TO SLAB - 553
I
A plate with deformed bar anchors is embedded in the edge of the slab. A bar is
welded to the edges of the plate. If there is no topping and the slab or flange is thick
enough, the connections can be recessed and caulked as shown.
I
Features: ...Quick. easy erection.
. ..Properly spaced connections distribute loading and transfer
I
diaphragm forces.
. ..Helps even out differential camber.
I
Disadvantages: ...Bar may not fit if joint is too wide.

Evaluation: . ..lA I
I
I
I
I
I
I
I
I
I
I
I
I
4.72
I
I
I SLAB TO SLAB - SS4

I A plate with reinforcing bars welded at 45’ angles is recessed slightly and
embedded in the slab. This provides the proper reinforcement cover and keeps the
connection from projecting above the top of the slab. A loose plate is welded at each
I end.
...Quick. easy erection.
Features:

I . ..Properly spaced connections distribute


diaphragm forces.
. ..Helps even out differential camber.
loading and transfer

I Disadvantages: . ..More material per connection than SS5.

Evaluation: . ..lC
I
I
I
I
I
I
I
I
I
I
I
I
4.73
I
I
SLAB To SLAB - SS5
I
A narrow plate with deformed bar anchors is embedded at about a 20’ angle. A
short piece of reinforcing bar is welded to the steel plates. I
Features: . ..Quick. easy erection.
. ..Properly spaced connections distribute
diaphragm forces.
loading and transfer I
...Helps even out differential camber.

Disadvantages: . ..Deformed bar anchors must be welded on at proper angle.


I
. ..If joint is too narrow, the bar may project above slab. If too wide,
bar may not stay in place.
I
Evaluation: . ..ZC

I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I 4.8 Wall to Foundation - WF

I
I
I
I
I
I
I
I
I
I
4.75
I
I
WALL TO FOUNDATION - WPl I
The wall panel is set on shims under each stem located at the centroid. Loose
angles are welded to plates embedded in the wall panel and foundation. Generally, two
connections per panel are provided. The space under the wall is usually filled with a dry-
I
pack grout.
Features: . ..Quick. easy erection.
I
...Pew tolerance problems.
Disadvantages: ...When placed below grade, may be awkward for welding. I
Evaluation: . ..lA
I
I
I
I
I
I
I
I
I
I
I
I
I
4.76
I
I
I WALL TO FOUNDATION - WP2

I Instead of being welded to a plate in the foundation, the angle is bolted; using
drilled-in expansion bolts. This eliminates the need for any hardware to be accurately
located in the foundation.

I The space under the wall is usually filled with a dry-pack grout.

Features: . ..Good corrective solution for missing plates.


I ...Easy field adjustment.

. ..Expansion bolts not as reliable for uplift.


Disadvantages:

I Evaluation: . ..lA

I
1
I
I
I
I
I
I
I
I
I
I
I
I
WALLTOFOUNDATION-WWF3
I
As with WFl and WF2, the wall panel is set on shims under each stem located at
the centroid. A loose angle is welded to plates embedded in the panel and foundation.
I
The bottom leg of the angle is usually turned under the stem when the footing isn’t wide
enough.
I
Features: ...Quick. easy erection.
. ..Allows access to connection when non-bearing panels are to be
removed and reused. I
Disadvantages: . ..Connection is away from main bcdy of panel (especially critical for
shear walls).
. ..Exterior connection is susceptible to moisture damage.
I
Evaluation: . ..2c
I
I
I
I
I
I
I
I
I
I
I
I
4.78
I
I
I WALL TO FOUNDATION - WF4

I This connection is used when the wall panel sets on a grade beam or wall. The wall
panel is set on shims under each stem located as near the centroid as possible. Loose
plates are welded to embedded plates on the face of the foundation wall and bolted into

I inserts in the wall panel. The gap under the wall is then filled with a dry-pack grout.

Features: . ..Quick. easy erection.

I Disadvantages: ...Connection doesn’t have good adjustment for tolerances normal to


the plane of the wall.
...Connection may be exposed if top of foundation is near finished
I Evaluation:
floor.

. ..2c

I
I
I
I
I
I
I
I
I
I
I
I
4.79
I
I
WALL TO FOUNDATION - WP5
I
This connection is used when cantilever moments must be resisted. The connection
at the base of the precast panel can be any of those shown in WFl - WF4. Coil rods are
threaded into inserts and cast into the floor slabs. Inset shows how strand lifting loops
I
can be used in conjunction with bent rebars to accomplish tie to floor slab.
Features: . ..Cantilever moment resistance provided.
I
Disadvantages: . ..Insert location (vertically) is critical.
...Slab construction requires care to assure that the bars are properly I
embedded.
Evaluation: ...2AS I
I
I
I
I
I
I
I
I
I
I
I
I
4.80
I
I
I WALLTOPOUNDATION-WWP6
I The foundation is cast with a slot to receive the wall panel. The wall panel is set
on shims as required and the shear key, which is 2 to 4 inches wider than the wall, is
filled in with grout. No other connection is made.
I Features: . ..Quick. easy erection.
. ..No tolerance problems.
I . ..Shear resistance perpendicular to wall.

. ..No uplift on overturning resistance other than dead load.


Disadvantages:

I Evaluation:
...No connection for wall panel during erection.

. ..lA

I
:
I ,,. .I...:”..+sa:..‘.
...:‘., . : :,
:..
I : :‘,
.,:i:_
.;;: :
,.,,&:‘:‘:’
9‘.
I
I
I
I
I
I
I
I
I
I 4.81
I
I
I
I
I
I
I
I
4.9 Slab to Wall - SW

I
I
I
I
I
I
I
I
I
I
4.82 I
I
I SLABTOWALL-SW1

I Precast slabs are set on high density plastic or hardboard bearing strips; leaving a
2 - 3 inch gap, end to end. If the wall is concrete masonry as shown, the cores are filled
in the last 2 - 3 courses and rebars are embedded at approximately 32” ox. Longitudinal
I reinforcement is added in the joint to tie the connection together. A composite topping,
reinforced with welded wire fabric, is poured and the next level of walI is constructed.

The connection is also applicable with a cast-in-place concrete wall.


I Features: . ..Quick. easy erection.
...Few tolerance problems.
I . ..Diaphragm forces transferred into wall.
...Topping poured in a large area.

I Disadvantages: . ..Good weather required to pour concrete topping which may delay
wall construction.

Evaluation: . ..lA
I
I
I
I
I
I
I
I
I
I
I
I
I
Precast slabs are erected on
SLAB TO WALL - SW2
high density plastic or hardboard bearing strips, simi-
I
lar to SWl; but leaving only about 1 inch tolerance gap, end to end. A bond beam is
shown as an alternate to the filled
level wall is constructed and later
cores in SWl. No vertical dowels are used. The next
a composite topping is poured; locking in the base of
I
the wall.

Features: . ..Quick. easy erection.


. ..Few tolerance problems.
I
Disadvantages:
. ..Bad weather doesn’t hold up construction
. ..Topping poured in smaller areas.
because of topping pour.
I
. ..No positive anchorage into wall.
. ..Little or no redundancy or ductility
limited to low rise buildings.
without tie steel. Therefore
I
Evaluation: . ..lCS
I
I
I
I
I
I
I
I
I
I
I
4.84 I
I
I
I SLABTOWALL-SW3

I This connection is used when a composite topping is not provided. Precast slabs are
erected the same as for SWI. Vertical reinforcement can be dowels as in SW1 or
continuous bars as shown here. Reinforcing bars are grouted into the keyways between
I slabs to provide structural integrity. Longitudinal bars assist in tying the structure
together and act as shear friction reinforcement for the slab grout keys.

I This connection is also applicable

Features:
with a cast-in-place

. ..Quick. easy erection.


concrete wall.

...Few tolerance problems.


I . ..Diaphragm forces transferred into wall.

Disadvantages: ...Space over the wall tends to become conjested with reinforcement.
I Evaluation: . ..lA

I
I
I
I
I
I
I
I
I
I
I
4.85
I
I
SLAB To WALL - SW4 I
Precast slabs are set on a continuous steel angle which is welded to embedded
plates in the wall. Some positive connection should be provided. Several alternates are
as illustrated in SB2, SB5 or SB6. Production restrictions or other conditions as discussed
I
for each, may determine the connection used.
This connection is generally only good for light loads. I
Features: ...Quick. easy erection.
. ..Wall can be continuous.
. ..No vertical tolerance problems.
I
Disadvantages: ...Connection may require fireproofing.
. ..Difficult to hold angle in place while welding.
I
. ..Little adjustment for tolerances normal to plane of wall, especially

Evaluation: . ..lC
when used with narrow wall panels.
I
I
I
I
I
I
I
I
I
I
I
4.86
I
I
I
I SLAB TO WALL - SW5

I This connection is usad when slabs cantilever over the wall. Dowels are grouted in
or embedded in the wall below at each slab joint. Precast slabs, with a notch at each
dowel location are erected on high density plastic or hardboard bearing strips. The
I notches are grouted with the keyways.

The dowels extend above the slab if there is a wall above or stop short of the top,

I as shown, if there is not.

Features: ...Quick. easy erection.


. ..Positive anchorage to wall.
I . ..Concealed connection.

Disadvantages: . ..Dowels may not align with notches.


I Evaluation:
...Notch is difficult with voided slabs.

. ..lCS

I
I
I
I
I
I
I
I
I
I
I
4.87
I
I
SLABTOWALL-SW6
I
The wall panel has either a continuous reinforced concrete corbel or a “buttorP
haunch, as shown in the inset. The tee sits on a bearing pad (typically neoprene or
I
laminated fabric) and has a top connection usually over each stem which can take a
variety of configurations as discussed in BCl.

Features:
I
. ..Quick. easy erection.
. ..Braces wall panel.
...Connection is protected when roofing is placed or topping is poured. I
Disadvantages: . ..Special forming required for corbel.
Evaluation: . ..lA
I
I
I
I
I
I
I
I
I
I
I
I
I
4.88
I
I
I SLABMWALL-SW7

I This connection is used for the special situation when the double tee frames into
the stem side of a precast wall panel. The double tee rests on a bearing pad on a weld-
on steel haunch attached to a plate in the stem. A top connection, similar to that of SW6

I is also required.

Note that the stem of the double tee is blocked back and the flange is notched
around the stem of the wall panel.
I Features: ...Relatively easy erection.
. ..Bolves special condition without overly complicating casting of units.
I Disadvantages:
. ..Braces wall panel.
. ..Connaction is exposed and may require protection.
. ..Tolerances are critical because of stresses caused by eccentricities.
I Evaluation: . ..2cs

I
I
I
I
I
I
I
I
I
I
I
4.89
I
I
SLABTOWALL-SW8
I
This connection may be used when roof slabs cantilever over the wall panels. Here,
the stem of the wall panel is blocked back and the flange notched instead of the double I
tee. The double tee rests on a bearing pad on the top of the stems and cantilevers out.
A plate cast in the bottom of the stem of the roof member is welded to the angle at the
support. This connection should only be used when the wall is flexible enough to avoid
volume change restraint problems.
I
Features: . ..No haunch is required.
. ..Provides desired cantilever.
I
. ..Braces wall panel.
Disadvantages: . ..Overhead weld is difficult; especially on high walls. I
. ..Connection may require fireproofing.
Evaluation: . ..2c I
I
I
I
I
I
I
I
I
I
I
I
4.90
I
I
I SLABTO WALL-SW9

I A loose angle with a vertical slot is bolted into an insert in the panel and welded to
a plate in the slab. The slot takes care of vertical adjustments made necessary by
camber or fabrication or erection tolerances. In the case of a roof, it also allows the
I slab to move with temperature changes.

The inset illustrates a welded alternate. Experience and calculations have shown

I that sufficient welded connections (typical spacing in most cases) can resist movements
due to temperature changes with no adverse effects.

Features: . ..Quick. easy erection.


I . ..No tolerance problems.
. ..No temperature change problems.
. ..Positive horizontal force transfer between slab and wall.
I Disadvantages: . ..Bolt sometimes hangs up in slot.

Evaluation: ...lC
I
I
I
I
I
I
I
I
I
I
I
4.91
I
I
SLABTOWALL-SW10
I
This is a variation of connection SW1 for use with stemmed deck units. Precast
tees are set on neoprene bearing pads with the flanges blocked back to the face of the
I
wall. If the wall is concrete masonry as shown, the cores are filled in the last 2 - 3
courses and rebars are placed at approximately 32” OX. Longitudinal reinforcement is
added in the joint to tie the connection together. A composite topping, reinforced with
welded wire fabric, is poured along with the connection and the next level of wall is
I
constructed.

The connection is also applicable with a cast-in-place concrete wall.


I
Features: . ..Diaphragm forces transferred into wall.
...Topping poured in large area. I
Disadvantages: . ..Good weather required to pour connection and topping.
...Area between stems must be formed. I
Evaluation: . ..2c
I
I
I
I
I
I
I
I
I
I
I
4.92
I
SLABTO WALL-SW11
This is a variation of connection SW10 when there is no composite topping. Bars
protruding from the flanges are bent down into the poured strip. The dowels stop short
of the top if there is no upper level. The rest of the connection is the same as SWIO.

Features: ...No tolerance problems.


. ..Diaphragm forces transferred into wall.
...Smooth surface for roofing.

Disadvantages: . ..Good weather required to pour connection.


...Area between stems must be formed.
I
Evaluation: . ..2C

I
I
I
I
I
I
I
I
I
I
I
I
4.93
I
I
I
SLABTOWALL-SW12 I
A common connection of a roof to a masonry wall. The last course is usually a
bond beam containing reinforcement. The flanges of the tees are not cut back and the
area between the stems can be left open, infilled The
withtees
masonry or closed end
are connected off to
withend as
I
precast, asbestos cement or translucent panels.
shown.
I
Design conditions such as significant shear force transfer may dictate the need for
an additional connection to the wall.
Features: ...No tolerance problems.
I
. ..No forming and pouring of connection required.

Disadvantages: ...No positive connection to wall. I


Evaluation: . ..lC
I
I
I
I
I
I
I
I
I
I
I
4.94 I
I
I SLABTOWALL-SW13

I This is a variation of connection SW4for stemmed units. The tees are set on a
continuous steel angle welded to embedded plates in the wall. Stiffeners are usually
required in the angle at the stem locations. An alternate to the continuous angle is
shown in the inset.
I A top connection is required as in SW6and discussed in BCl.

I Features: . ..Quick. easy erection.


. ..Wall can be continuous.
. ..Usually no vertical tolerance problems.
I Disadvantages: . ..Connection may require protection.
. ..Connection is recommended for light loads only.
. ..Little adjustment for tolerances normal to plane of wall, especially
I when used with narrow wall panels.
Evaluation: . ..2c
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
4.10 Beam to Wall - BW I
I
I
I
I
I
I
I
I
I
I
4.96
I
I
I BEAMTOWALL-BWl
Similar to SW6. For heavier beam loads, design of wall must consider the
I distribution of the loads.

I Evaluation: . ..lA

I
I
I
I
I
I
I
I
I
,
I

4.97
I
BEAMTOWALL-BW2 I
A pocket is cast into the wall to receive the beam. The beam sits on a bearing pad
and the pocket is later grouted, preferably after dead loads are in place. A top connec-
tion, as discussed in BCl could also be provided. Axial shortening of beam due to volume
I
change should be considered when designing depth of recess.
This connection @ more commonly used with cast-in-place walls, but can also be
I
used with precast panels.
Features: . ..Minimum of embedded hardware. I
. ..Clean. concealed connections.
Disadvantages: . ..Pocket dimensions must be planned so that beam can “swing” into
place-especially critical if this connection is used at both ends.
I
Evaluation: ...lC I
I
I
I
I
I
I
I
I
I
I
I
4.98
I
I BEAMTOWALL-BW3

I A dowel is placed in a bond beam or grouted masonry cores. See BC2 for other
features.

Evaluation: . ..2c
I

1
I

4.99
I
BEAM TO WALL - BW4
I
This connection is used when the beam is required to provide lateral restraint to
the wall. The bearing plate is installed on a mortar bed immediately after the concrete I
is placed in the bond beam.

Features: . ..Positive tie to wall. I


Disadvantages: . ..Requires close coordination of trades.
...Level placement of plate difficult.
. ..Volume change shortening may damage the wall.
I
Evaluation: . ..2cs
I
I
I
I
I
I
I
I
I
I
I
I
I
4.100
I
I
I
I
I
I
I
I
I 4.11 Wall to Wall - WW

I
I
I
I
I
I
I
I
I
I
I 4.101
I
WALLTO WALL--WI
I
A plate is cast into the upper wall, anchored with deformed stud bars which lap
with vertical reinforcement in the panel. A threaded rod or coil rod projects from the
I
lower panel and bolts through an oversized hole in the plate. This rod also laps with
vertical reinforcement to provide a continuous vertical tension tie through the system.
The formed pocket is later patched to conceal the connection. Proper vertical elevation I
is obtained with shims or a leveling nut under the plate.

Features: . ..Minimum amount of hardware required.


...No welding.
I
. ..Connection is concealed and protected after patching.

Disadvantages: . ..Requires accurate placement of projecting bolt.


I
. ..Projecting bolt subject to damage in handling.

Evaluation: . ..lA I
I
I
I
I
I
I
I
I
I
I
I
I
I
I WALL TO WALL - WW2

I A section of structural steel tube is east into the upper panel. Continuous steel
rods which project from the lower panel are bolted to the top and bottom of the tube.
With high strength bars and a heavy steel tube, a high ultimate resistance to uplift can be
I attained.

Elevation control can be accomplished by leveling nuts as shown, or with shims.


I Features: . ..High ultimate tensile strength.
...No welding.

I ...Connection is concealed and protected


. ..Continuous vertical connection.
after patching.

Disadvantages: ...Requires accurate placement and alignment of hardware.


I . ..Hardware is expensive.

Evaluation: . ..2C
I
I
I
I
I
I
I
I
I
I
I
I 4.103
I
WALLTO WALL--W3
I
The upper wall is cast with an embedded plate in a recessed pocket. The lower wall
is cast with a plate projecting from the top which fits up against and is welded to the
I
plate in the upper wall. The pocket is later patched to conceal the connection. During
erection, the panel is shimmed to the proper elevation and later dry packed.
I
The lower plate fits in the joint between the slab and the notched wall or between
the two slabs at a typical interior connection. The reinforcement on the plates can be
lapped with vertical reinforcement in the panels, providing continuity through the
connection.
I
Features: ...Positive connection between walls.
. ..Connection is concealed and protected from fire after patching.
I
. ..Continuous vertical reinforcement through the connection.
Disadvantages: . ..Plate alignment in the wall is critical. I
Evaluation: ...2c

I
I
I
I
I
I
I
I
I
I
I
I
I WALLMWALL-WW4

I The sleeve connectors shown here receive reinforcing bars from each end and are
later filled with an expansive grout to form a continuous connection. The device is
proprietary and tests have shown they are capable of developing the full strength of the
I bar. (See also CC%.)

The sleeve connector can be placed in any of positions shown, but placement in the

I upper panel is usually preferred.

Features: . ..Continuity through the connections.


. ..Connection is concealed and protected.
I Disadvantages: . ..No connection between walls until splice sleeves are grouted.
. ..Sleeve connectors and sleeve grout are proprietary.
I Evaluation:
. ..Hardware placement is critical.

. ..2A

I
I
I
I
I
I
I
I
I
I
I
4.105
I
I
WALLTO WALL--W5
I
This connection uses vertical post-tensioning to connect the walls. Sleeves are cast
in the wall panels at the plant. In most applications, the tensioning is done one floor at a
I
time. The tendons (usually bars) are attached to an anchor (at the bottom floor) or a
coupler (at intermediate floors). The panels are then “threaded” over the bars. The bars
are tensioned and anchored, leaving enough projection to attach a coupler to receive the
bar for the next level.
I
Features: . ..Fully continuous tension tie.
. ..Connection is totally concealed and protected.
I
...High tensile strength.
. ..Post-tensioning reduces drift in high rise buildings.
I
Disadvantages: ...Complex erection procedure.
. ..Post-tensioning is an added operation.
. ..Alignment of sleeves is critical.
. ..Hardware is expensive.
I
Evaluation: . ..lC I
I
I
I
I
I
I
I
I
I
I
4.106
I
I
I WALLTOWALL-WW6

I Plates are cast in the wall panel and anchored with studs and/or welded reinforcing
bars. A loose plate is welded across the joint. In the corner condition shown in the inset,
a loose steel angle is used instead of the loose plate.
I Features: ...Ample adjustment allowance.
. ..When recessed, connection is concealed and protected.

I Disadvantages:
...Good shear transfer.

. ..Rigid connection.
...Possible volume change problems.
I Evaluation: . ..lA

I
I
I
I
I
I
I
I
I
I
I
I
I 4.107
I
WALLTOWALL-WW'I
I
Inserts or bolts welded to steel plates are cast into the panels. The loose plate has
slots in opposite direction on each side to allow both vertical and horizontal adjustment.
I
If properly placed, the slots will also allow some movement for volume changes. If a
rigid connection is desired, the plates can be later welded.
. ..Quick erection.
I
Features:
. ..Volume change strain relief.
. ..When recessed, connection is concealed and protected.
...No welding required.
I
Disadvantages: . ..Less field adjustment than WW6.
...Shear transfer between panels unreliable without welding.
I
Evaluation: . ..lA
I
I
I
I
I
I
I
I
I
I
I
I
4.108
I
I
I
WALLTOWALL-WW8
I angle.
Steel angles are cast into the panels. A bar spans the joint and is welded to each

I Features: ...Inexpensive hardware.

Disadvantages: . ..Requires close joint tolerances.


I ...If reinforcing bar used, weldability
. ..Rigid. unyielding connection.
may be questionable.

I Evaluation: . ..2c

I
I
I
I
I
I
I
I
I
I
I
I
I 4.109
I
I
REFERENCES - PART 4

I
4.1 “Connection Details for Precast-Prestressed Concrete Buildings”, Prestressed
Concrete Institute, 1963.
I
4.2 “Connection Details for Precast, Prestressed Concrete”, by the Australian
Prestressed Concrete Group. Published by Cement and Concrete Association of

4.3
Australia about 1965.

“Precast Concrete Connection Details-Structural Design Manual”, Stupre-


I
4.4
Society for Studies on the use of Precast Concrete, Netherlands, 1978.

“PC1 Design Handbook, Precast and Prestressed Concrete”, Second Edition.


I
Prestressed Conocrete Institute, Chicago, IL, 1978.

4.5 “PC1 Manual on Design of Connectins for Precast, Prestressed Concrete”, First
I
Edition. PC& Chicago, IL 1973.

4.6 “PC1 Manual for Structural Design of Architectural Precast Concrete”,


Edition. Prestressed Concrete Institute, Chicago, IL 1977.
First I
4.7 Cazaly, L., and Huggins, M., “Canadian Prestressed Concrete Institute Handbook”, I
Canadian Prestressed Concrete Institute, 1964.

I
I
I
I
I
I
I
I
I
4.110
I
I METRIC CONVERSION

I SI Base units

I Quantity

length
Name

meter
Symbol

Ill
mass kilogram kg

I electric
time
current
second
ampere
s
A
thermodynamic temperature
I amount
luminous
of substance
intensity
mole ,llOl

I SI Supplementary Units

I
I SI Derived Units

I Quantity Name Symbol


In Terms
of
Other Units
In Terms
of
Base Units

I frequency hertz HZ - 5-1

force newton N - m l kg l s2
pressure, stress Pascal Pa N/m2 ,,-I. kg. s-2

I en$g;;,uk,

power
quantity
jOUk

watt
J
w
Nom
J/s
m’
m2
l

l
kg
kg
l

l
s-*
s3

I SI Derived Units

In Terms In Terms

I Quantity Description of
Other Units
Of
Base Units

I area square meter - m’


volume cubic meter - ma
density, mass density kilogram per cubic - kg/m3

I specific volume
meter
cubic meter par
kilogram
- ml/kg

I moment of force
heat capacity
specific heat capacity
newton meter
joule per kelvin
joule par kilogram
N-m
J/K
J/kg-K
mr.kg*s-z
,,,‘.kg.s-Z.K-’

,,,‘.s-2.K-1
kelvin
I thermal conductivity watt per meter
kelvin
Wlm*K m*kg*s-3 K-l

I
I
METRIC CONVERSION I
I
Quantity
Other Units to Use with SI

Name Symbol Value in SI Units


I
minute min 1 min = 60s
I
hour h 1 h = 3600s
plane angle degree
minute
0 1’ = (a/180) rad
1 ’ = (~110 800) rad
I
temperature

mass
degree
Celsius
tonne
“C

t
1 “C (interval) = 1 K
0°C = 273.15 K
1 t = 1000 kg
I
I
Conversion Factors I
U.S. Customary to SI

To convert from to multiply by I


millimeter (mm)
I
I
I
I
I
I
pound/square foot kilopascal (kPa)
pound/square inch (psi) kilopascal (ItPa) I
I
I
I
I METRIC CONVERSION

I
To convert from to multiply by

I Bending Moment or Torque

I inch-pound
foot-pound
newton-meter
newton-meter
(Nom)
(N*m)
0.1130
1~.356
foot-kip newton-meter (Nom) 1356

I
I MXS
I

I pound (avdp)
ton (short, 2000 lb)
kilogram
kilogram
(kg)
(kg)
0.4536
907.2
ton (short, 2000 lb) tonne (t) i 0.9072

I I

Mass per Volume


I pound/cubic foot kilogram/cubic meter (kg/m’) 16.02

I pound/cubic yard kilogram/cubic meter (kg/m’) 0.5933

I Temperature

deg Fahrenheit (F) deg Celsius (C) C = (F - 32) 519

I deg Fahrenheit (F), interval deg Celsius (C), interval 519

I Other

I section
moment
modulus
of inertia
in.l
in:
mm3
mm’
16,387
416,231
Coefficient of heat transfer,

I Btulft’ hr ‘F
Modulus of elasticity, psi
W/m’* “C
MPa
5.678
0.006895
Thermal conductivity,

I Btu/in./ft’ hr “F
Thermal expansion in.lin. ‘F
W/m* “C
mm/mm* “C
0.1442
1.800
fc = mf’,, psi MPa 0.063036 J

I
I
I
-
I
I
I

I
.-
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
Prestressed Concrete Institute
201 North Wells Street
I
Chicago, Illinois 60606
Telephone 3121346-4071 I
Printedin U.S.R
I

You might also like