Вы находитесь на странице: 1из 0

printed on June 24, 2003

Lecture - 1-1
LECTURE 1 - BACKGROUND AND PHILOSOPHY
1.1 OBJECTIVE OF THE LESSON
The objective of this lesson is to acquaint the student with the
historical background surrounding the development of the LRFD
Specification. The material presented will provide insight into how the
basic decisions regarding the technical basis of the specification were
arrived at and how the organization to actually write the specification
was developed. The basic concepts behind the probabilistic reliability-
based calibration will be introduced.
1.2 HISTORICAL DEVELOPMENT
1.2.1 Background
The apparent start of the process leading to the LRFD
Specification was the initiation of NCHRP Project 20-7/31, entitled
"Development of Comprehensive Bridge Specification and
Commentary", in August of 1986. In reality, the process leading to this
decision had started almost ten years earlier.
In the late 1970's, the Ontario Ministry of Transportation and
Communication, now known as the Ministry of Transportation, decided
to develop its own bridge design specification, rather than continue
utilizing the AASHTO Standard Specification for Highway Bridges. In
the process of considering the basis for this new Specification, a
decision was taken to base it on probabilistic limit states. A Code
Control Committee, chaired by Mr. Paul F. Csagoly, P. Eng., began to
develop background material on the variability of loads and the
components that make up resistance, including basic variabilities,
such as the dispersion of the values for yield strength of metals,
compressive strength of concrete and the variation of sizes in factory-
made and field-made products. A major study to determine the
statistical variation in vehicle weights and configurations was also
completed. During the same time frame, the basic process for
calculating the statistical reliability of a bridge component, based on
the mean values of the applied loads and the parameters that went
into the determination of resistance, and the standard deviations of
these values was also developed. A process for determining a
combination of multipliers on load and resistance to achieve a level of
reliability, to be further explained in Article 1.3, was also developed.
In 1979, the first edition of the Ontario Highway Bridge Design
Code (OHBDC) was released to the design community as North
America's first calibrated, reliability-based limit state specification.
Since that time, the OHBDC has been updated in 1983 and 1993 and
re-released. Very significantly, the code contained a companion
volume of commentary.
printed on June 24, 2003
Lecture - 1-2
As more and more U. S. engineers became familiar with the
OHBDC, they recognized a certain logic in the calibrated limit states
design and began to question whether the AASHTO Specification
should be based on a comparable philosophy of determining the
safety of structures. Many research projects, undertaken by the
NCHRP, the National Science Foundation (NSF), and various states
were bringing new information on bridge design faster than it could be
critically reviewed and, where appropriate, adopted into the AASHTO
Specification. It was also becoming clear that the many revisions
which had occurred to the AASHTO Specification had resulted in
numerous inconsistencies and the appearance of a patchwork
document.
In the Spring of 1986, a group of State Bridge Engineers or
their representatives met in Denver and drafted a letter to the
Subcommittee on Bridges and Structures indicating their concern that
the AASHTO Specification was falling behind the times. They also
raised the concern that the Technical Committee structure, operating
under the Subcommittee on Bridges and Structures, was not able to
keep up with emerging technologies. Presentations were made at two
regional meetings to mixed reception. Nonetheless, this group,
identified below, planted the seed which led to the development of the
LRFD Specification.
Name 1986 Affiliation
James E. Roberts California Department of Transportation
H. Henrie Henson Colorado Department of Highways
Paul F. Csagoly Florida Department of Transportation
Ho Lum Wong Michigan Department of Transportation
Charles S. Gloyd Washington Department of Transportation
In July of 1986, a group of State Bridge Engineers met with the
staff of the NCHRP to consider whether a project could be developed
to explore the points raised in the Denver letter. This led to NCHRP
project 20-7/31 "Development of Comprehensive Bridge Specifications
and Commentary", a pilot study conducted by Modjeski and Masters,
Inc. with Dr. John M. Kulicki as Principal Investigator. The following
is a list of tasks for this project:
Task 1 - Review of the philosophy of safety and coverage
provided by other specifications.
Task 2 - Review AASHTO documents, other than the Standard
Specification, for their potential for inclusion into a standard
specification.
Task 3 - Assess the feasibility of a probability-based
specification.
Task 4 - Prepare an outline for a revised AASHTO
Specification for Highway Bridge Design, commentary, and
printed on June 24, 2003
Lecture - 1-3
present a proposed organizational process for completing such
a document.
The review of other specifications and the trends in the development
of new specifications included a review of work done in Canada
(especially the Province of Ontario), Great Britain, the Federal
Republic of Germany and Japan. Personal contacts with practitioners
and researchers in other countries provided information on the
emerging directions of specification development and insight into what
designers were doing to implement specifications, and, in some cases,
what designers were choosing not to implement based on a
perception of unnecessary complication. Information collected from
these various sources indicated that most of the First World countries
appeared to be moving in the direction of a calibrated, reliability-
based, limit states specification.
Task 2 can best be summarized as a search for gaps and
inconsistencies in the 13
th
Edition of the AASHTO Standard
Specifications for Highway Bridges. "Gaps" were areas where
coverage was missing; "inconsistencies" were internal conflicts, or
contradictions of wording or philosophy. Many gaps and
inconsistencies were found and they are summarized in the list shown
in Table 1.2.1-1.
TABLE 1.2.1-1 GAPS AND INCONSISTENCIES
GENERAL FORMAT
Division II
Commentary
Presentation Format
ANALYSIS AND DESIGN PHILOSOPHY
Use of more Refined Design Methods for
Girder Bridges
Improved Slab Design
Effective Flange Width
Bridge Dynamics
Foundation Design Methods
The Current LFD Provisions in the
AASHTO Specifications
Curved Girder Bridges
PERFORMANCE OF MEMBERS AND
SYSTEMS
Modern Bearing Systems
Features of Prestressed Concrete Design
Fatigue of Prestressed Girders
Shielded or Blanketed Strands
Design of Compression Members
Partial Prestressing
Prestress Losses
Local Stress Requirements
Time Dependent Concrete Properties
Foundation Design for Lateral Loads
Compression Plate Design
Anchorage Zone Stresses
printed on June 24, 2003
Lecture - 1-4
The Effect of Skew
ADDITIONAL LOADS
The Live Load Model
Thermal and other Environmental Loads
Ship Collision
Erection Engineering and Construction
Loads
Combination of Load
TYPES OF CONSTRUCTION NOT
COVERED OR PARTIALLY COVERED
Segmental Concrete Bridges
Cable-Stayed Bridges
Multi-Web Box Girder Bridges
Design for Shear and Torsion in Concrete
Members by Space Truss Analogy
United Treatment of Concrete Design
Continuity Joints for Prestressed I-Beams
Made Continuous for Live Load
Horizontal Shear Requirements and
Composite Sections
Features of Steel Design
Carrying Capacity of Distinctly
Unsymmetric Plate Girders
Splices in Overdesigned Members
Net Section Requirements for Built-up
Members
K Factors for Compression Members
Friction Joints
Riveted Construction
Sealing Requirements
Deflection Criteria
Metal Deck Systems
Proprietary Wall Systems
Details which are Sensitive to Distortion-
Induced Fatigue
Connection Design
The BS5400 Fatigue Detail Catalog
With respect to Task 3 and the feasibility of using probability-
based limit states design, a review of the philosophy used in a variety
of specifications resulted in three possibilities, two of which are
already included in the current specification. They are:
Allowable stress design which treats each load on the
structure as equal from the view point of statistical variability.
A "common sense" approach may be taken to recognize that
some combinations of loading are less likely to occur than
others, e.g., a load combination involving a 160 km per hour
wind, dead load, full shrinkage and temperature may be
thought to be far less likely than a load combination involving
printed on June 24, 2003
Lecture - 1-5
the dead load and the full design live load. For example, in the
13
th
Edition and others, the former load combination was
permitted to produce a stress equal to four-thirds of the latter.
Load Factor Design - In which a preliminary effort was made
to recognize that the live load, in particular, was more highly
variable than the dead load. This thought is embodied in the
concept of using a different multiplier on dead and live load,
e.g., a load combination involving 130% of the dead load
combined with the 217% of the live load, and requiring that a
measure of resistance based primarily on the estimated peak
resistance of a cross-section exceed the combined load.
Reliability-based design which seeks to take into account
directly the statistical mean resistance, the statistical mean
loads, the nominal, or notional, value of resistance, the
nominal or notional value of the loads and the dispersion of
resistance and loads as measured by either the standard
deviation or the coefficient of variation, i.e, the standard
deviation divided by the mean. This process can be used
directly to compute probability of failure for a given set of
loads, statistical data and the Designer's estimate of the
nominal resistance of the component being designed. Thus,
it is possible to vary the nominal resistance to achieve a
criteria which might be expressed in terms such as the
component (or system) must have a probability of failure of
less than 0.0001, or whatever variable is acceptable to society.
Alternatively, the process can be used to target a quantity
known as the "reliability index" which is somewhat, but not
directly, relatable to the probability of failure. Based on this
"reliability index", it is possible to reverse engineer a
combination of load and resistance factors to achieve a
specific reliability index. This is discussed in Article 1.3.
While some specifications are being developed in terms of the
"probability of failure", it was generally agreed that it would be more
appropriate to use the reliability index process to develop load and
resistance factors. That way, design could proceed in a process
directly analogous to load factor design as it appeared in the 13
th
Edition of the Standard Specifications.
In May of 1987, the findings of NCHRP Project 20-7/31 were
presented to the AASHTO Subcommittee on Bridges and Structures
outlining the information above and indicating that seven options
appeared to be available for consideration. They were:
Option 1 - Keep the Status Quo
Option 2 - Table Consideration of LRFD for the Short-Term
Option 3 - Immediate Adoption of the OHBDC
printed on June 24, 2003
Lecture - 1-6
Option 4 - Replace Current Specification with LRFD
Immediately
Option 5 - Replace Current LFD with LRFD in the near term
Option 6 - Develop LRFD for evaluation only
Option 7 - Develop LRFD as a Guide Specification
A recommendation was made to proceed to:
develop a probability-based limit states specification,
fill as many of the gaps and inconsistencies as possible, and
develop a commentary to the specification.
Under the direction of then Chairman Robert Cassano of California,
the Subcommittee directed the NCHRP to develop a project to
complete this task. This led to NCHRP Project 12-33 which was also
entitled "Development of Comprehensive Specification and
Commentary", which was started in July of 1988, by Modjeski and
Masters, Inc.
1.2.2 Organization of Project
1.2.2.1 RESEARCH TEAM
A hierarchial structure was established consisting of a Principal
Investigator and Co-Principal Investigator from Modjeski and Masters,
Inc., a Code Coordinating Committee, as identified in Table 1.2.2.1-1,
and 15 working groups called task groups also identified in Table
1.2.2.1-1. Additionally, an Editorial Committee was developed in
charge of the responsibility of assembling the information and making
it editorially and technically consistent.
The original plan was to have the Code Coordinating
Committee meet on a regular basis and to judicate the technical
content of the Specification. While the Code Coordinating Committee
met several times in the early part of the development of the
specification, it became apparent that to meet the schedule imposed
on the project, the Editorial Committee would have to deal directly with
the Task Group Chairman.
TABLE 1.2.2.1-1 - NCHRP 12-33 PROJECT TEAM
John M. Kulicki, Principal Investigator
Dennis R. Mertz, Co-Principal Investigator
Scott A. Sabol, Program Officer (1992-1993)
Ian M. Friedland, Senior Program Officer (1988-1993)
printed on June 24, 2003
Lecture - 1-7
CODE COORDI NATI NG
COMMITTEE
John M. Kulicki, Chairman
-Modjeski and Masters, Inc.
John J. Ahlskog
-FHWA
Richard M. Barker
-Virginia Polytechnic Institute
and State University
Robert C. Cassano
-Imbsen & Associates, Inc.
Paul F. Csagoly
-SRD Engineering, Inc.
James M. Duncan
-Virginia Polytechnic Institute
and State University
Dennis R. Mertz
-University of Delaware
Theodore V. Galambos
-University of Minnesota
Andrzej S. Nowak
-University of Michigan
Frank D. Sears
-Consultant
NCHRP PANEL
Veldo M. Goins, Chairman
-Oklahoma DOT
Roger Dorton
-Buckland and Taylor Ltd
Steven J. Fenves
-Carnegie-Mellon University
Richard S. Fountain
-Parsons, Brinckerhoff, Quade
and Douglas, Inc.
C. Stewart Gloyd
-Corridor Design Management
Group
Stanley Gordon
-FHWA
Geerhard Haaijer
-AISC
Clellon L. Loveall
-Tennessee DOT
Basile Rabbat
-Portland Cement Assoc. of
Research & Development
James E. Roberts
-California DOT
Arunprakash M. Shirole
-New York State DOT
James T. P. Yao
-Texas A & M University
Luis Ybanez
-Texas State DOT (Retired)
EDITORIAL COMMITTEE
*
John M. Kulicki, Chairman
Paul F. Csagoly
Dennis R. Mertz
Frank D. Sears
*
with appreciation to Modjeski
and Masters, Inc.'s staff
members:
Diane M. Long
Scott R. Eshenaur
Chad M. Clancy
Robert P. Barrett
David M. Barrett
Donald T. Price
Nancy E. Kauhl
Malden B. Whipple
Wagdy G. Wassef
Raymond H. Rowand
Charles H. Johnson
TASK GROUPS
General Design Features
Frank D. Sears, Chairman
Stanley R. Davis
-
Ivan M. Viest
-Consultant
Loads and Load Factors
Paul F. Csagoly, Chairman
Peter G. Buckland
-Buckland and Taylor Ltd.
Eugene Buth
-Texas A & M University
James Cooper
-FHWA
C. Allin Cornell
-Stanford University
James H. Gates
-CALTRANS
Michael A. Knott
Greiner, Inc.
Fred Moses
-University of Pittsburgh
Andrzej S. Nowak
Robert Scanlan
-Johns Hopkins University
Analysis and Evaluation
Paul F. Csagoly, Chairman
Peter Buckland
Ian G. Buckle
-State University at Buffalo
Roy A. Imbsen
-Imbsen & Associates, Inc.
Jay A. Puckett
-University of Wyoming
Wallace W. Sanders, Jr.
-Iowa State University
Frieder Seible
-University of California - San
Diego
William H. Walker
-University of Illinois at
Urbana-Champaign
Concrete Structures Steel Structures Aluminum Structures
printed on June 24, 2003
Lecture - 1-8
Robert C. Cassano, Chairman
John H. Clark
-Anderson, Bjornstad, Kane
and Jacobs, Inc.
Michael P. Collins
-University of Toronto
Paul F. Csagoly
David P. Gustafson
-CRSI
Antoine E. Naaman
-University of Michigan
Paul Zia
-North Carolina State
University
Don W. Alden
-Imbsen & Associates, Inc.
Frank D. Sears, Chairman
John Barsom
-USS Division of USX Corp.
Karl Frank
-University of Texas at Austin
Wei Hsiong
-Consultant
William McGuire
-Cornell University
Dennis R. Mertz
Roy L. Mion
-AISC Marketing,Inc.
Charles G. Schilling
- Consultant
Ivan M. Viest
Michael A.Grubb
-AISC Marketing, Inc.
Frank D. Sears, Chairman
Teoman Pekoz
-Cornell University
Wood Structures
Andrzej S. Nowak, Chairman
Baidar Bakht
-Ministry of Transp. of Ontario
R. Michael Caldwell
-TBTA
Donald J. Flemming
-Minnesota DOT
Hota V. S. Gangarao
-West Virginia University
Joseph F. Murphy
-Structural Reliability
Consultants
Michael A. Ritter
-USDA Forest Products Lab
Raymond Taylor
-Ministry of Transp. of Ontario
Thomas G. Williamson
-American Institute of Timber
Construction
Deck Systems
Paul F. Csagoly, Chairman
Barrington deVere Batchelor
-Queens University
Daniel H. Copeland
-Grid Manufacturer's Assoc.
Gene R. Gilmore
-IKG/Greulich
Richard E. Klingner
-University of Texas at Austin
Roman Wolchuk
-Consultant
Foundations
J. Mi chael Duncan, Co-
Chairman
Ri chard M. Barker, Co-
Chairman
Walls, Piers and Abutments
J. Mi chael Duncan, Co-
Chairman
Ri chard M. Barker, Co-
Chairman
James Withiam
-D'Appolonia
Buried Structures
James Withiam
Bridge Railings
Ralph W. Bishop, Chairman
-California DOT
Eugene Buth
James H. Hatton, Jr.
-FHWA
Teddy J. Hirsch
-Texas A & M University
Robert A. Pege
- New Jersey DOT
J o i n t s , Be a r i n g s a n d
Accessories
Charles W. Purkiss, Chairman
-California DOT
Ian G. Buckle
Earthquake Provisions Advisory
Group
Ian Buckle, Chairman
Robert Cassano
James Cooper
Calibration
Andrzej S. Nowak, Chairman
C. Allin Cornell
Dan M. Frangopol
printed on June 24, 2003
Lecture - 1-9
John J. Panak
-Texas State Dept. of
Highways
and Public Transportation
David Pope
-Wyoming DOT
Charles W. Roeder
-University of Washington
John F. Stanton
-NYSBA
James Gates
Roy Imbsen
Geoffrey Martin
-University of Southern
California
-University of Colorado
Theodore V. Galambos
Roger Green
-University of Waterloo
Fred Moses
Kamal B. Rojiani
-VPISU
1.2.3 Project Schedule
The original plan called for three drafts, which were released
and reviewed as follows:
The first draft was released in April of 1990 and was totally
uncalibrated. The primary intent was to show coverage and
organization. This draft was released to the AASHTO Bridge
Engineers, the FHWA, all members of the NCHRP Panel and
Task Group Members, and several private authorities. All told,
it was reviewed by about 250 engineers, because many of the
Department of Transportations circulated it to in-house
experts. Approximately 4,000 comments were received
concerning the first draft, all were read and reviewed, and
many were discussed with Task Group Chairmen or sent
directly to them. Many of the comments were included in the
second draft, but there was no written response to the
questions.
A second draft was released in late April of 1991 to the same
group of people. Additionally, it was noted at several regional
and national conferences on bridge engineering that all
interested parties could obtain a copy of the draft specification
at their cost, and that they would be free to submit review
comments. This second draft contained a preliminary set of
load and resistance factors which changed relatively little in
subsequent drafts. Approximately 6,000 comments were
received for this draft and were processed as outlined above.
The third draft was submitted in April of 1992 and was
reviewed in the same process that was used for the second
draft. For this draft, about 2,000 comments were received and
they were processed as described above.
After reviewing the third draft, the NCHRP Panel determined
that the specification was approaching a draft that could be considered
for a ballot item, but that additional work would be worthwhile and
would reduce modifications needed in the future. Accordingly, the
project was extended to include a fourth draft, whose scope included
the following items:
Continue to review the distribution factors developed under
NCHRP Project 12-23 and which were included in the
proposed specification,
printed on June 24, 2003
Lecture - 1-10
Continue to refine calibration,
Consider further the need for special short-span live loads,
Further refine and verify the proposed strip width method for
calculating moments and deck slabs,
Develop an index to the specification,
Convert to the SI system of measurement,
Develop further trial designs, and
Complete the text for fourth draft.
This fourth draft was submitted in March of 1993 and was
accepted as a ballot item at the May 1993 meeting of the
Subcommittee of Bridge and Structures.
One of the most valuable features of the process of developing
this specification was two rounds of trial designs. In 1991, and again
in 1992, various States and industry groups volunteered to do
comparative designs using the 14
th
edition of the Standard
Specifications and the LRFD Specifications. Additionally, interested
industry groups also organized their own series of trial designs and
contributed information and critiques based on that work. Fourteen
States and several industry groups participated in the initial 1991
designs, and 22 States and several industry groups worked on the
1992 set. As would be expected, the 1992 set were more complete
and included nine slab bridges, 20 concrete beam bridges, 9 steel
beam and girder bridges, 1 truss, 1 segmental concrete bridge, 2
wood bridges and 5 culverts, and a series of retaining wall designs.
The designs included substructure, superstructure and pile and spread
footing foundations. Additionally, a comprehensive set of prestressed
beam bridges were evaluated by industry and contributed to the
project.
These two series of trial designs achieved several important
objectives:
They exposed areas where further development of load
models and resistance formulations was necessary, and where
further calibration was advisable.
They demonstrated that the specification, though considerably
longer and more comprehensive than the Standard
Specification, was nonetheless readable and workable.
They pointed to numerous areas where improvements and
clarification in the wording could be made.
printed on June 24, 2003
Lecture - 1-11
They vastly broadened the base of practicing engineers who
were becoming conversant with the LRFD Specification.
All things considered, the two trial design sessions, which
required supplementary meetings of the Subcommittee on Bridges
and Structures, proved to be one of the most important steps in the
development and adoption of the LRFD Specification.
1.2.4 Project Objectives
There were several objectives in the development of this new
specification. They may be summarized as:
To develop a technically state-of-art specification which would
put U. S. practice at or near the leading edge of bridge design.
To make the specification as comprehensive as possible and
include new developments in structural forms, methods of
analysis and models of resistance.
To the extent consistent with the thoughts above, keep the
specification readable and easy to use, bearing in mind that
there is a broad spectrum of people and organizations involved
in bridge designs.
To keep specification-type wording and not to develop a
textbook.
To encourage a multi-disciplinary approach to bridge design,
particularly in the area of hydraulics and scour, foundation
design and bridge siting.
To place increasing importance on the redundancy and
ductility of structures.
Many changes had to be made in the content and appearance
of the Standard Specification to achieve the objectives outlined above.
Areas of major changes are identified below:
The introduction of a new safer philosophy of safety - LRFD.
The identification of four limit states to be discussed later.
The development of new load factors.
The development of new resistance factors.
The relationship of the chosen reliability level, the load and
resistance factors, and load models through the process of
calibration.
The development of improved load models necessary to
achieve adequate calibration, including a new live load model.
printed on June 24, 2003
Lecture - 1-12
Revised techniques for analysis and the calculation of load
distribution.
A combined presentation of plain, reinforced and prestressed
concrete.
The introduction of limit state-based provisions for foundation
design and soil mechanics.
Expanded coverage on hydraulics and scour.
Changes to the earthquake provisions to eliminate the seismic
performance category concept by making the method of
analysis a function of the importance of the structure.
Inclusion of large portions of the Guide Specification for
Segmental Concrete Bridge Design.
Inclusion of large portions of the FHWA Specification for ship
collision.
Expanded coverage on bridge rails based on crash testing,
with the inclusion of methods of analysis for designing the
crash specimen.
The introduction of the isotropic deck design process.
The development of a parallel commentary.
It was the underlying principle of NCHRP 12-33 to make as
much use of existing research findings as possible. The project was
not supposed to involve the development of new information, although
some limited work was necessary to tie information together to make
a comprehensive specification.
Any effort to develop a Specification on the scale of the
AASHTO LRFD Specification for Highway Bridge Design has to reach
a point where upgrading the technical content for new ideas must be
stopped in order to finish the test and publish the document. This
does not stop the tide of new ideas. Future changes must be
expected. Some of the areas where continued development should
be expected and encouraged include:
Continue development of a database upon which to project
bridge loads. This is particularly true of live load for which it
was initially thought that much information would be
determined from weight-in-motion (WIM) studies. For various
reasons, much of the WIM data was not directly usable in the
development of the live load model. Coincident with NCHRP
12-33, an FHWA-sponsored project involving the
instrumentation of bridges throughout the Country and the
measurements of loads in correlation of analysis. This project
has generated large amounts of data, and continuing efforts
printed on June 24, 2003
Lecture - 1-13
should be made to extract from that data information for further
refinements of load models and analysis techniques.
There is a continuing need to refine and verify foundation
resistance and deformation. More work needs to be done on
the large scale testing of foundations, the determination of
group action, and the amount of movement which is
acceptable. This latter point was a subject of considerable
discussion during the development of NCHRP 12-33 as
Structural Engineers and Foundation Engineers are apt to view
this issue differently. Clearly, a multi-disciplinary consensus
is necessary.
The issue of temperature gradient was also hotly debated.
There appears to be general agreement that the temperature
gradient does exist, as has been demonstrated in many
studies. The real problem appears to be related to the
structural effect actually caused by the gradient and its joint
action with other loads.
Further simplification of load distribution is warranted, as is
extension to explicitly include unequal spans and splayed
girders. This could take the form of further refinement of
orthotropic plate models, which were considered in the
development of the LRFD specification, and suitable PC-type
computer program to do a detailed analysis.
The joint probability of load occurrence remains an issue of
much interest. How much live load should be applied with an
earthquake loading, should other loads be applied with that
same combination? How much more refinement of wind
loadings can be done before site-specific studies are
necessary? How should ice, wind and other loads be
combined? Should ship and vehicle collision be applied
simultaneously with scour and earthquake and other loadings?
Continued development of reliability theory should involve
more emphasis on system rather than component reliability,
the use of second order methods, improved methods of
projecting the all important "tails" of measured data, the
development of larger and more inclusive databases, and the
inclusion of aging and deterioration models.
Re-evaluate countrywide loadometer survey, future projections
of truck traffic, etc., and relate them to fatigue criteria.
Continue further refinement on temperature gradient, ice
loads, stream flow forces, debris forces, including development
of threshold design values below which there is no need to
apply these forces.
Finally, it is time to start developing a "North American
Highway Bridge Design Specification". The common LRFD
printed on June 24, 2003
Lecture - 1-14
(probabilistic limit states) philosophy utilized in Canadian and U. S.
Specifications should make such a joint specification, applicable to
Mexico if they agreed and helped to develop it, more attainable now
than at any time in the last 15 years. Such a multi-National
specification would be consistent with the Eurocode approach and the
concept of a North American Free Trade Zone.
1.3 SUMMARY OF RELIABILITY CONSIDERATION
1.3.1 Overview of a Probability-Based Specification
The investigation of probability-based limit states design
started on a note of some skepticism regarding whether this
philosophy was mature enough in its development to encompass the
combination of art and science involved in bridge engineering. After
considering the underlying principles of service load design, load
factor design and limit states design, it became apparent that of these
three possibilities, probability-based limit states design was
unquestionably the most comprehensive and rational way to proceed.
Finally, for clarity, it was decided that the AASHTO version of limit
states design would be termed Load and Resistance Factor Design
(LRFD), like the AISC Specification.
A consideration of probability-based reliability theory can be
simplified considerably by initially considering that natural phenomena
can be represented mathematically as normal random variables, as
indicated by the well-known bell-shaped curve, Figure 1.3.1-1. Use of
this assumption leads to closed form solutions for areas under parts
of this curve as given in Table 1.3.1-1, which can conveniently answer
the following questions:
What percentage of the total number of values fall within a
given range Y X Z? The answer to this question is given
by the area bounded by the two values Y and Z, as shown in
Figure 1.3.1-1.
What percentage of the total values are such that X Z? This
is shown by the shaded area in Figure 1.3.1-2.
The first question and its statistical ramifications are already included
in the AASHTO Standard Specifications for Highway Bridges in the
fatigue design provisions of Article 10.3.1, in which the allowable
fatigue stress ranges of the various categories were defined by the
"95% confidence" limits. To put this in the prospective of Figure 1.3.1-
1, this is equivalent to saying that 95% of all the details tested at a
given stress range failed at a number of cycles bounded by the values
of Y and Z. In this context, the value of Y and the value of Z are each
located two standard deviations on either side of the mean or average
value.
The question illustrated by Figure 1.3.1-2 deals very explicitly
with the problem of defining loads and resistance of members.
printed on June 24, 2003
Lecture - 1-15
z 0 1 2 3 4 5 6 7 8 9
0.0 0.0000 0.0040 0.0080 0.0120 0.0160 0.0199 0.0239 0.0279 0.0319 0.0359
0.1 0.0398 0.0438 0.0478 0.0517 0.0557 0.0596 0.0636 0.0675 0.0714 0.0754
0.2 0.0793 0.0832 0.0871 0.0910 0.0948 0.0987 0.1026 0.1064 0.1103 0.1141
0.3 0.1179 0.1217 0.1255 0.1293 0.1331 0.1368 0.1406 0.1443 0.1480 0.1517
0.4 0.1554 0.1591 0.1628 0.1664 0.1700 0.1736 0.1772 0.1808 0.1844 0.1879
0.5 0.1915 0.1950 0.1985 0.2019 0.2054 0.2088 0.2123 0.2157 0.2190 0.2224
0.6 0.2258 0.2291 0.2324 0.2357 0.2389 0.2422 0.2454 0.2486 0.2518 0.2549
0.7 0.2580 0.2612 0.2642 0.2673 0.2704 0.2734 0.2764 0.2794 0.2823 0.2852
0.8 0.2881 0.2910 0.2939 0.2967 0.2996 0.3023 0.3051 0.3078 0.3106 0.3133
0.9 0.3159 0.3186 0.3212 0.3238 0.3264 0.3289 0.3315 0.3340 0.3365 0.3389
1.0 0.3413 0.3438 0.3461 0.3485 0.3508 0.3531 0.3554 0.3577 0.3599 0.3621
1.1 0.3643 0.3665 0.3686 0.3708 0.3729 0.3749 0.3770 0.3790 0.3810 0.3830
1.2 0.3849 0.3869 0.3888 0.3907 0.3925 0.3944 0.3962 0.3980 0.3997 0.4015
1.3 0.4032 0.4049 0.4066 0.4082 0.4099 0.4115 0.4131 0.4147 0.4162 0.4177
1.4 0.4192 0.4207 0.4222 0.4236 0.4251 0.4265 0.4279 0.4292 0.4306 0.4319
1.5 0.4332 0.4345 0.4357 0.4370 0.4382 0.4394 0.4406 0.4418 0.4429 0.4441
1.6 0.4452 0.4463 0.4474 0.4484 0.4495 0.4505 0.4515 0.4525 0.4535 0.4545
1.7 0.4554 0.4564 0.4573 0.4582 0.4591 0.4599 0.4608 0.4616 0.4625 0.4633
1.8 0.4641 0.4649 0.4656 0.4664 0.4671 0.4678 0.4686 0.4693 0.4699 0.4706
1.9 0.4713 0.4719 0.4726 0.4732 0.4738 0.4744 0.4750 0.4756 0.4761 0.4767
2.0 0.4772 0.4778 0.4783 0.4788 0.4793 0.4798 0.4803 0.4808 0.4812 0.4817
2.1 0.4821 0.4826 0.4830 0.4834 0.4838 0.4842 0.4846 0.4850 0.4854 0.4857
2.2 0.4861 0.4864 0.4868 0.4871 0.4875 0.4878 0.4881 0.4884 0.4887 0.4890
2.3 0.4893 0.4896 0.4898 0.4901 0.4904 0.4906 0.4909 0.4911 0.4913 0.4916
2.4 0.4918 0.4920 0.4922 0.4925 0.4927 0.4929 0.4931 0.4932 0.4934 0.4936
2.5 0.4938 0.4940 0.4941 0.4943 0.4945 0.4946 0.4948 0.4949 0.4951 0.4952
2.6 0.4953 0.4955 0.4956 0.4957 0.4959 0.4960 0.4961 0.4962 0.4963 0.4964
2.7 0.4965 0.4966 0.4967 0.4968 0.4969 0.4970 0.4971 0.4972 0.4973 0.4971
2.8 0.4974 0.4975 0.4976 0.4977 0.4977 0.4978 0.4979 0.4979 0.4980 0.4981
2.9 0.4981 0.4982 0.4982 0.4983 0.4984 0.4984 0.4985 0.4985 0.4986 0.4986
3.0 0.4987 0.4987 0.4987 0.4988 0.4988 0.4989 0.4989 0.4989 0.4990 0.4990
3.1 0.4990 0.4991 0.4991 0.4991 0.4992 0.4992 0.4992 0.4992 0.4993 0.4993
3.2 0.4993 0.4993 0.4994 0.4994 0.4994 0.4994 0.4994 0.4995 0.4995 0.4995
3.3 0.4995 0.4995 0.4995 0.4996 0.4996 0.4996 0.4996 0.4996 0.4996 0.4997
3.4 0.4997 0.4997 0.4997 0.4997 0.4997 0.4997 0.4997 0.4997 0.4997 0.4998
3.5 0.4998 0.4998 0.4998 0.4998 0.4998 0.4998 0.4998 0.4998 0.4998 0.4999
3.6 0.4998 0.4998 0.4999 0.4999 0.4999 0.4999 0.4999 0.4999 0.4993 0.4999
3.7 0.4999 0.4999 0.4999 0.4999 0.4999 0.4999 0.4999 0.4999 0.4999 0.4999
3.8 0.4999 0.4999 0.4999 0.4999 0.4999 0.4999 0.4999 0.4999 0.4999 0.4999
3.9 0.5000 0.5000 0.5000 0.5000 0.5000 0.5000 0.5000 0.5000 0.5000 0.5000
Table 1.3.1-1 - Areas Under Portion of Normal Probability Density Curve
printed on June 24, 2003
Lecture - 1-16
Figure 1.3.1-1 - Normal Distribution Curve Showing Distribution
Bounded by the Values "Y" and "Z"
Figure 1.3.1-2 - Normal Distribution Curve Showing Portion of
Distribution Less than or Equal to "Z"
If we now accept the notion that both load and resistance are
normal random variables, we can plot the bell-shaped curve
corresponding to each of them in a combined presentation dealing
with distribution as the vertical axis against the value of load, Q, or
resistance, R, as shown in Figure 1.3.1-3. The mean value of load
and the mean value of resistance is also shown, as is a second value
somewhat offset from the mean value, which is the "nominal" value,
or the number that designers calculate the load or the resistance to
be. The ratio of the mean value divided by the nominal value is called
the "bias". The objective of a design philosophy based on reliability
theory, or probability theory, is to separate the distribution of
resistance from the distribution of load, such that the area of overlap,
i.e., the area where load is greater than resistance, is tolerably small,
say one in 10,000. The objective of a LRFD approach to a design
specification is to be able to define load factors, shown as
Qn
in
Figure 1.3.1-3, and resistance factors, shown as
Rn
in Figure 1.3.1-3,
in a way that forces the relationship between the resistance and load
to be such that the area of overlap is less than or equal to the value
that a code-writing body accepts. Note in Figure 1.3.1-3 that it is the
nominal load and the nominal resistance, not the mean values, which
are factored.
printed on June 24, 2003
Lecture - 1-17
Figure 1.3.1-3 - Separation of Loads and Resistance
The relationship between the nominal (design) load, mean load
and factored load is shown in Figure 1.3.1-4. The shaded area in
Figure 1.3.1-4 is equal to the probability of exceeding the factored load
value.
Figure 1.3.1-4 - Probability Density Function, f
x
(q), of Load
Component, Q
i
; Mean Load, Q
m
, Nominal (design) Load, Q
n
, and
Factored Load,
i
Q
n
(Nowak, 1993)
The relationship between the nominal (design) resistance,
mean resistance and factored resistance is shown in Figure 1.3.1-5.
The shaded area in Figure 1.3.1-5 is equal to the probability of
exceeding the factored resistance value.
Figure 1.3.1-5 - Probability Density Function, f
R
(X), of Resistance, R;
Mean Resistance, R
m
, Nominal (design) Resistance, R
n
, and Factored
Resistance, R
n
(Nowak, 1993)
printed on June 24, 2003
Lecture - 1-18
A conceptual distribution of resistance minus load, combining
the individual curves discussed above, is shown in Figure 1.3.1-6,
where the area of overlap from Figure 1.3.1-3 is shown as negative
values, i.e., those values to the left of the origin.
Figure 1.3.1-6 - Definition of Reliability Index,
It now becomes convenient to define the mean value of resistance
minus load as some number of standard deviations, , from the origin.
The variable is called the "reliability index". The problem with this
presentation is that the variation of the quantity, resistance minus load,
is not explicitly known. Much is already known about the variation of
loads by themselves or resistances by themselves, but the difference
between these has not yet been quantified. However, from probability
theory, it is known that if load and resistance are both normal and
random variables then the standard deviation of the difference is:
(1.3.1-1)

(R Q)

2
R

2
Q
Given the standard deviation, and considering Figure 1.3.1-6, we can
now define the reliability index, , as:
(1.3.1-2)

R Q

2
R

2
Q
Comparable closed-form equations can also be established for other
distributions of data, e.g., log-normal distribution. It is very important
to realize that a "trial and error" process is available for solving for
when the variable in question does not fit one of the already existing
closed-form solutions.
In a reliability-based code in the purest sense, the designer is
asked to calculate the value of provided by his design and then
compare that to a code-specified tolerable value. A designer would
require much knowledge of reliability theory to apply such a pure
reliability-based code. Alternatively, through a process of calibrating
printed on June 24, 2003
Lecture - 1-19
load and resistance factors by trial designs, it is possible to develop
load and resistance factors so that the design process looks very
much like the existing Load Factor Design Methodology.
The process of calibrating load and resistance factors starts
with Equation 1.3.1-2 and the basic design relationship; the factored
resistance must be greater than or equal to the sum of the factored
loads:
(1.3.1-3)
R Q
i
x
i
Solving for the average value of resistance yields:
(1.3.1-4)
R Q
2
R

2
Q
R
1


i
x
i
Using the definition of bias, indicated by the symbol , Equation 1.3.1-
4, leads to the second equality in Equation 1.3.1-4. A straightforward
solution for the resistance factor, is:
(1.3.1-5)


i
x
i
Q
2
R

2
Q
Unfortunately, Equation 1.3.1-5 contains three unknowns, i.e., the
resistance factor, , the reliability index, , and the load factors, .
The acceptable value of the reliability index, , must be chosen
by a code-writing body. While not explicitly correct, we can conceive
of as an indicator of the fraction of times that a design criteria will be
met or exceeded during the design life, analogous to using standard
deviation as an indication of the total amount of population included or
not included by a normal distribution curve. Utilizing this analogy, a
of 2.0 corresponds to approximately 97.3% of the values being
included under the bell-shaped curve, or 2.7 of 100 values not
included. When is increased to 3.5 for example, now only two
values in approximately 10,000 are not included. It is more accurate
to think of as a basis for comparing the relative safety of populations
structures.
Assuming that a code-writing body has established a value of
, Equation 1.3.1-5 still indicates that both the load and resistance
factors must be found. One way to deal with this problem is to select
the load factors which apply to all materials and structural action,
assume trial resistance factors for each material and action being
considered and then calculate the reliability indicies. This process has
been used by several code-writing authorities.
The steps in the process is as follows:
printed on June 24, 2003
Lecture - 1-20
Factored loads can be defined as the average value of load,
plus some number of standard deviation of the load, as shown
as the first part of Equation 1.3.1-6 below.
(1.3.1-6)
i
x
i
x
i
n
i
x
i
nV
i
x
i
Defining the "variance", V
i
, as being equal to the standard
deviation divided by the average value, leads to the second
half of Equation 1.3.1-6. Utilizing the concept of bias one more
time, Equation 1.3.1-6 can now be condensed into Equation
1.3.1-7.
(1.3.1-7)

i
1 nV
i
Thus, it can be seen that load factors can be written in terms
of the bias and the variance, as depicted in Figure 1.3.1-7.
This gives rise to the philosophical concept that load factors
can be defined so that all loads have the same probability of
being exceeded during the design life. This is not to say that
the load factors are identical, just that the probability of the
loads being exceeded is the same.
Figure 1.3.1-7 - Graphical Presentation of Equation 1.3.1-7 (Nowak
and Lind, 1979)
Using Equation 1.3.1-5, for a given set of load factors and trial
resistance factors, the value of the reliability index can be
calculated for various types of structural members and for
various load components, e.g., shear, moment, etc. on the
various structural components. Computer simulations of a
representative body of structural members can be done,
yielding a large number of values for the resistance factor.
Reliability indices are then grouped by structural member and
by load component to determine if they cluster around the
printed on June 24, 2003
Lecture - 1-21
target reliability index. If close clustering results, a suitable
combination of load and resistance factors has been obtained.
If close clustering does not result, a new trial set of load and/or
resistance factors can be used and the process repeated until
the reliability indices do cluster around the target.
The resulting load and resistance factors taken together will
yield reliability indices close to the target value selected by the
code-writing body as acceptable.
The process above appears to be rather illusive. Fortunately,
other jurisdictions had utilized this calibration process and found it to
yield reasonable load and resistance factors, which was also the case
in NCHRP 12-33. Figure 1.3.1-8 shows the dispersion of the reliability
indices observed for bridges designed to the AASHTO Standard
Specifications within the Province of Ontario. After defining load and
resistance factors through the process outlined above, analysis of
these same set of bridges produced reliability indices clustered around
the target value of 3.5, as shown in Figure 1.3.1-9. The reason that
the values do not plot exactly on the horizontal straightline, indicated
by a reliability index of 3.5, is that the resistance factors did not cluster
at exactly the same number. Thus, when reasonable and
conservative interpretations of the resistance factor values are utilized,
the reliability indices will generally be above the target value and an
unavoidable amount of scatter will still result. As can be seen in
Figure 1.3.1-9, a handful of the comparative values were significantly
below the target reliability index, and this indicated that additional
design provisions were necessary for this particular group of bridges.
printed on June 24, 2003
Lecture - 1-22
Figure 1.3.1-8 - Range of Reliability Indices Obtained Using AASHTO
Standard Specifications (Nowak and Lind, 1979)
Figure 1.3.1-9 - Range of Reliability Indices Obtained with Reliability-
Based OHBDC (Nowak and Lind, 1979)
The chief advantages for a probability-based LRFD
specification are as follows:
A more uniform level of safety throughout the system will
result.
That measure of safety will be a function of the variability of
loads and resistance.
printed on June 24, 2003
Lecture - 1-23
Designers will have an estimate of the probability of meeting
or exceeding the design criteria during the design life.
The potential exists to place all structural materials and
methods of construction on equal footing.
A realistic rational framework for future development of the
specification will be available.
Proponents of future changes in materials and construction
techniques will be asked to provide the same measure of
reliability that all current materials and construction methods
will be asked to meet.
Designers will have better understanding of where and how
uncertainties of load and resistance models are accounted for,
and will be able to relate past performance.
There are certainly some disadvantages in basing a specification on
this philosophy. Increased design effort will certainly result, as it is
realistic to expect that a greater number of load and resistance factors
will be available. However, the designer will need little or no
knowledge of reliability theory. There will be start-up costs in re-
education and in the upgrade of design aids and design software. In
summary, when considering the benefits, the obstacles did not seem
to justify staying with the status quo.
The probability-based LRFD for bridge design was seen as a
logical extension of the current Load Factor Design procedure. The
Service Load Design does not recognize that various loads are more
variable than others. The introduction of the Load Factor Design
methodology brought with it the major philosophical change of
recognizing that some loads are more accurately represented than
others. The conversion to probability-based LRFD methodology could
be thought of as a mechanism to more systematically and rationally
select the load and resistance factors than was done with the
information available when Load Factor Design was introduced.
Consideration was given to the impact that probability-based
LRFD would have on the design community. It was expressly not
intended that a specification be developed that requires all design
engineers to be well-grounded in probability or reliability theory.
Rather, a specification was developed by this process that does not
appear to be a radical departure from the current Load Factor Design
provisions. As previously stated, there are more individual load and
resistance factors, but the process of computing a combined load
group is basically similar to Load Factor Design. Likewise, the
process of computing the resistance of a structural member are based
on principles not radically different from those in the Standard
Specifications, although, where possible, the design requirements are
upgraded to the current state-of-the-art.
printed on June 24, 2003
Lecture - 1-24
1.4 OVERVIEW OF THE CALIBRATION PROCESS
1.4.1 Outline of the Calibration Process
The following steps are the major phases of the calibration of
the load and resistance factors for the LRFD Specification:
Develop a database of sample current bridges
Extract load effects by percentage of total load
Develop a simulation bridge set for calculation purposes
Estimate the reliability indices implicit in current designs
Revise loads-per-component to be consistent with the LRFD
Specification
Assume load factors
Vary resistance factors until suitable reliability indices result as
described in Article 1.3.1
The outline above assumes that suitable load factors are assumed.
If the process of varying the resistance factors and calculating the
reliability indices does not converge to a suitable narrowly grouped set
of reliability indices, then the load factor assumptions must be revised.
In fact, several sets of proposed load factors were investigated to
determine their effect on the clustering of reliability indices.
1.4.2 Development of a Sample Bridge Database
Approximately 200 representative bridges were selected from
various regions of the United States by requesting sample bridge
plans from various states. The selection was based on structural-type
material and geographic location to represent a full-range of materials
and design practices as they vary around the Country. Anticipated
future trends were also considered by questionnaires sent to selected
states. One-hundred and seven sets of plans were received from
which the 200 representative bridges were selected. Obviously, some
plan sets contained more than one bridge or the bridge contained
several separable units. The list of structures provided by the State
Departments of Transportation is given in Table 1.4.2-1.
For bridges selected from within this database, moments and
shears were calculated for the dead load components, the live load
and the dynamic load allowance. Nominal or design values were
calculated using the 1989 edition of the AASHTO Standard
Specifications. The statistically projected live load and the notional
values of live load force effects were calculated. Resistance was
calculated in terms of moment and shear capacity. For each structure,
both the nominal design resistance, indicated by the cross-section
shown on the plans, and the minimum actual required resistance,
according to the 1989 AASHTO, were developed. Generally
speaking, the nominal resistance is larger than the minimum value,
printed on June 24, 2003
Lecture - 1-25
because standard available sizes of plates and rebar and other
components that comprise resistance do not exactly meet the
theoretical requirement. Additionally, a phenomenon known as the
"designers bias" is implicit in actual designs. This factor results from
the tendency of designers to "bump up" a given component to the next
available commercial size.
printed on June 24, 2003
Lecture - 1-26
Table 1.4.2-1 - Selected Bridges
STRUCTURAL TYPE REQUESTED SPAN (m) PROVIDED SPAN (m)/STATE
Steel, Simple Span
Rolled Beams, Non-composite 12 to 25 14.6 - PA
18.0 - MI
25.3 - PA
Rolled Beams, Composite 15 to 25 14.6 - PA
14.9 - PA
15.2 - PA
15.5 - PA
20.4 - PA
23.2 - PA
24.4 - PA
26.2 - PA
Plate Girder, Non-composite 30 to 45 23.8 - PA
30.5 - PA
Plate Girder, Composite 30 to 55 31.4 - MI
33.2 - PA
37.2 - MI
Box Girder 30 to 55 None
Through-truss 90 to 120 91.4 - PA
92.4 - PA
94.8 - PA
121.0 - PA
Deck Truss 60 to 120 61.0 - NY
76.2 - NY
91.4 - NY
121.9 - NY
Pony Truss 45 30.5 - OK
31.4 - PA
91.4 - PA
Arch 90 to 150 109.7 - NY
132.9 - NY
192.0 - NY
222.5 - NY
Tied Arch 90 to 180 163.1 - NY
Steel, Continuous Span
Rolled Beams 15-20-15 to 25-30-25 22.6-18.3 - PA
25.9-24.4-25.9 - MI
23.2-29.3-24.4-18.3 - PA
Plate Girder 30-36-30 57.9-54.9 - MI
36.6-45.7-36.6 - MI
61.0-61.0-61.0 - KY
91.4-91.4-91.4-91.4 - KY
59.4-59.4-59.4-59.4 - KY
61.0-61.0-61.0-61.0-61.0-61.0 - KY
Box Girder 30-36-30 to 90-120-90 31.4-31.4-31.4 - MD
37.5-37.5-37.5 - MD
43.3-45.7-31.4 - MD
37.2-49.4-37.2 - IL
35.4-42.1-42.1-42.1-35.4 - IL
45.7-50.9-53.3-53.3-50.9-45.7 - IL
Through-truss 120 None
Deck Truss 120 None
Tied Arch 90-150 None
Reinforced Concrete, Simple Span
printed on June 24, 2003
STRUCTURAL TYPE REQUESTED SPAN (m) PROVIDED SPAN (m)/STATE
Lecture - 1-27
Slab 6 to 12 9.1 - OK
T-beam 12 to 24 12.2 - IL
12.2 - OK
13.1 - IL
15.2-15.2 - OK
18.3 - IL
Arch-barrel 12 None
Arch-rib 18 None
Reinforced Concrete, Continuous Span
Slab, Two-span 9-9
12-12
None
Slab, Three-span 8-8-8 None
Solid Frame 12 12.2 - CA
T-beam, Frame 16 None
T-beam, Two-span 15-15
21-21
18.9-18.9 - CO
21.6-21.6 - CO
T-beam, Three-span 12-15-12 to 15-21-15 11.6-15.2-11.6 - TN
12.2-15.5-12.2 - TN
12.7-15.5-12.2 - TN
14.0-17.1-11.9 - TN
14.3-19.8-14.3 - TN
16.2-22.3-16.2 - TN
15.2-21.6-12.8 - TN
Arch None
Box, Three-span 18-24-18 to 23-27-23 21.0-36.3-29.3 - MD
Prestressed Concrete, Simple Span
Slab 9 to 12 None
Voided Slab 9 to 15 None
Double T 12 to 18 11.9 - CO
Closed Box CIP 38 None
AASHTO beam 15 to 30 23.2 - MI
23.2 - CO
31.1 - TX
31.1 - PA
32.0 - PA
31.4 - MI
33.5 - CO
36.0 - TX
36.0 - CO
39.6 - TX
42.1 - CO
Bulb 18 to 36 None
Box Girder 24 to 36 22.6 - PA
22.6 - PA
25.0 - CA
29.0 - CA
31.1 - CA
31.7 - CA
35.4 - CA
36.0 - CA
36.0 - CA
38.1 - CA
38.1 - CA
Prestressed Concrete, Continuous Span
printed on June 24, 2003
STRUCTURAL TYPE REQUESTED SPAN (m) PROVIDED SPAN (m)/STATE
Lecture - 1-28
Slab 10-10 to 12-15-12 None
Voided Slab 15-21-15 to 32-32 None
AASHTO Beam 25 to 33 None
Post-tensioned AASHTO Beam 30-30 None
Bulb None
Box 19.8-19.8 - CA
26.5-25.9 - CA
28.3-26.2 - CA
31.4-31.1 - CA
32.6-31.1 - CA
33.5-48.8 - CA
36.0-30.8 - CA
61.0-61.0 - CA
18.3-24.4-18.3 - CA
21.0-25.0-18.0 - CA
22.9-27.4-22.9 - CA
21.0-28.0-21.0 - CA
23.2-27.4-23.2 - CA
21.6-25.9-21.6 - CA
20.1-25.9-15.8 - CA
Wood
Saw Beam 5.5 - MN
Glulam Beam - Nailed 14.9-15.2-14.9 - MN
Glulam Beam - Dowelled None
Glulam Beam - Composite None
Truss 15.2-30.5-30.5-14.9 - MN
Arch None
Deck - Nailed 9.8-9.8-9.8 - MN
Deck - Composite None
Deck - Prestressed Transversely 13.4 - MN
Deck - Prestressed Longitudinally
1.4.3 Extraction of Load Effects
For each of the bridges in the database, the load indicated by
the contract drawings was subdivided by the following characteristic
components:
The dead load due to the weight of factory-made components
The dead weight of cast-in-place components
The dead weight due to asphaltic wearing surfaces where
applicable
The dead weight due to miscellaneous items
The live load due to the HS20 loading
The dynamic load allowance or impact prescribed in the 1989
AASHTO Specification
printed on June 24, 2003
Lecture - 1-29
Full tabulations for all these loads for the full-set of bridges in the
database are presented in Nowak, 1993.
In summary, the combination of the tandem with the uniform
load and the HS20 with the uniform load, were shown to be an
adequate basis for a notional design load in the LRFD Specification.
1.4.4 Development of the Simulated Bridge Set
Based on the relative amounts of the loads identified in the
preceding article for each of the combination of span and spacing and
type of construction indicated by the database, a simulated set of 175
bridges was developed which was comprised of:
Twenty-five non-composite steel girder bridge simulations for
bending moment with spans of 9, 18, 27, 36 and 60 m, and for
each of those spans, spacings of 1.2, 1.8, 2.4, 3.0 and 3.6 m.
Representative composite steel girder bridges for bending
moments having the same parameters as those identified
above.
Representative reinforced concrete T-beam bridges for
bending moments having spans of 9, 18, 27 and 39, with
spacings of 1.2, 1.8, 2.4 and 3.6 m in each span group.
Representative prestressed concrete bridges for moments
having the same span and spacing parameters as those used
for the steel bridges.
Representative steel girder bridges for shear having the same
span and spacing parameters as those identified for bending
moment.
Representative reinforcing concrete T-beams for shear having
the same span and spacing parameters indicated previously
for bending moment.
Representative prestressed concrete girder bridges for shear
having the same span and spacing parameters as previously
indicated for prestressed beams.
Full tabulations of these bridges and their representative amounts of
the various loads are presented in Nowak, 1993.
1.4.5 Calculated Reliability Indices and Selection of Target Value
The reliability indices were calculated for each simulated and
each actual bridge for both shear and moment. The range of reliability
indices which resulted from this phase of the calibration process is
presented in Figure 1.4.5-1. It can be seen that a wide range of
values were obtained using the current specifications, but this was
printed on June 24, 2003
Lecture - 1-30
anticipated based on previous calibration work done for the Ontario
Highway Bridge Design Code (OHBDC).
The most important parameters which determine the reliability
index for beam and girder bridges are the girder spacing and the span
length. In general, reliability indices are higher for larger girder
spacing, due to the conservatism of the S/N-type distribution factors
utilized for conventional beam and girder bridges in the 1989 AASHTO
Specification.
Figure 1.4.5-1 - Reliability Indices Inherent in the 1989 AASHTO
Standard Specification
These calculated reliability indices, as well as past calibration
of other specifications, serve as a basis for the selection of the target
reliability index,
T
. A target reliability index of 3.5 was selected for the
OHBDC and is under consideration for other reliability-based
specifications. A consideration of the data shown in Figure 1.4.5-1
indicates that a of 3.5 is indicative of past practice. Hence, this
value was selected as a target for the new calibration.
1.4.6 Load and Resistance Factors
1.4.6.1 LOAD FACTORS
The parameters of bridge load components are summarized
in Table 1.4.6.1-1. These data and Equation 1.3.1-7 enable an initial
set of trial load factors to be calculated in a rational way.
printed on June 24, 2003
Lecture - 1-31
Table 1.4.6.1-1 - Parameters of Bridge Load Components
LOAD COMPONENT
BIAS
FACTOR
COEFFICIENT
OF
VARIATION
Dead Load, D
1
1.03 0.08
Dead Load, D
2
1.05 0.10
Dead Load, D
3
1.00 0.25
Live Load (with impact) 1.10-1.20 0.18
Various sets of load factors, corresponding to different values
of n, are presented in Table 1.4.6.1-2. The relationship is also shown
in Figure 1.4.6.1-2.
Table 1.4.6.1-2 - Considered Sets of Load Factors
LOAD
COMPONENT n = 1.5 n = 2.0 n = 2.5
Dead Load, D
1
1.15 1.20 1.24
Dead Load, D
2
1.20 1.25 1.30
Dead Load, D
3
1.375 1.50 1.65
Live Load (with
impact)
1.40-1.50 1.50-1.60 1.60-
1.70
Figure 1.4.6.1-2 - Load Factors vs. n (Nowak, 1993)
Recommended values of load factors correspond to n = 2. For
simplicity of the designer, one factor is specified for D
1
and D
2
, =
printed on June 24, 2003
Lecture - 1-32
1.25. For D
3
, weight of asphalt, = 1.50. For live load and impact, the
value of load factor corresponding to n = 2 is = 1.60. However, a
more conservative value of = 1.75 is utilized in the LRFD code.
1.4.6.2 RESISTANCE FACTORS
The acceptance criterion in the selection of resistance factors
is how close the calculated reliability indices are to the target value of
the reliability index,
T
. Various sets of resistance factors, , are
considered. Resistance factors used in the code are rounded off to
the nearest 0.05. For each value of , the minimum required
resistance, R
LRFD
, is determined from the following equation:
(1.4.6.2-1)
R
LRFD
1.25D 1.50 D
A
1.75 (L I)

where D is dead load, except of D


A
, which is the weight of the asphalt
surface. The load factors are equal to the recommended values from
Section 1.4.6.1.
The calculations were performed using the load components
for each of the 175 simulated bridges. For a given resistance factor,
material, span and girder spacing, a value of R
LRFD
is calculated using
Equation 1.4.6.2-1. Then, for each value of R
LRFD
and corresponding
loads, the reliability index is computed. For easier comparison with
1989 AASHTO, a resistance ratio, r, is defined as
(1.4.6.2-2)
r
R
LRFD
R
HS20
Resistance ratio is an indication of the actual change of the code
requirements. Value of r > 1 corresponds to LRFD code being more
conservative than 1989 AASHTO, and r < 1 corresponds to LRFD
being less conservative than 1989 AASHTO.
Values of r and were calculated for live load factors, = 1.75.
For comparison, the results are also shown for live load factor, =
1.60. The calculations are performed for the resistance factors, ,
listed in Table 1.4.6.2-1.
printed on June 24, 2003
Lecture - 1-33
Table 1.4.6.2-1 - Considered Resistance Factors
MATERIAL LIMIT
STATE
RESISTANCE
FACTORS,
LOWER UPPER
Non-Composite Steel Moment 0.95 1.00
Shear 0.95 1.00
Composite Steel
Moment 0.95 1.00
Shear 0.95 1.00
Reinforced Concrete
Moment 0.85 0.90
Shear 0.90 0.90
Prestressed Concrete
Moment 0.95 1.00
Shear 0.90 0.95
1.4.6.3 RECOMMENDED LOAD AND RESISTANCE FACTORS
The recommended load factors are listed in Table 1.4.6.3-1
and recommended resistance factors are given in Table 1.4.6.3-2.
Table 1.4.6.3-1 - Recommended Load Factors
LOAD COMPONENT LOAD FACTOR,
Dead Load (except asphalt overlay) 1.25
Dead Load (asphalt overlay and
utilities)
1.50
Live Load (including impact) 1.75
printed on June 24, 2003
Lecture - 1-34
Table 1.4.6.3-2 - Recommended Resistance Factors
MATERIAL
LIMIT
STATE
RESISTANCE
FACTOR,
Non-Composite Steel
Moment 1.00
Shear 1.00
Composite Steel
Moment 1.00
Shear 1.00
Reinforced Concrete
Moment 0.90
Shear 0.90
Prestressed Concrete
Moment 1.00
Shear 0.90
Reliability indices were recalculated for each of the 175
simulated cases and each of the actual bridges from which the
simulated bridges were produced. The range of values obtained using
the new load and resistance factors is indicated in Figure 1.4.6.3-1.
Comparing the values of reliability indices obtained with the
1989 AASHTO Specification and the LRFD Specification indicates that
a considerable improvement in the clustering of reliability index values
has been obtained. This is a direct result of the integration of the load
factor, resistance factor, accurate load models and suitable resistance
models. NCHRP 12-33 was not charged to make a wholesale re-
adjustment of the inherent safety in the highway system. Selection of
the target reliability index of 3.5 is consistent with that view. However,
a fully consistent philosophy has been established for the
specification, as indicated by the tightly clustered reliability index value
shown in Figure 1.4.6.3-1.
printed on June 24, 2003
Lecture - 1-35
Figure 1.4.6.3-1 - Reliability Indices Inherent in LRFD Specification
At any future time, AASHTO may decide that more or less safety
(reliability) is desired. Should such a decision be made, a consistent
means is now available to adjust load factors and resistance factors
to achieve any increment in reliability.
One of the early concerns about the development of a
probability-based LRFD Specification was that it would be used as a
basis for reducing the strength of bridges. A comparison was made
of the apparent resistance demands required by the 1989 AASHTO
Specification and the LRFD Specification. For purposes of
comparison, the "demand" is taken as a sum of the factored loads
divided by a resistance factor. Such a comparison is shown in Figure
1.4.6.3-2 for the simulated bridges used in the calibration process,
based on the second draft. Some small variations are possible due to
refinements made as the Specification developed. This figure
indicates that generally slightly more structure will be required based
on the factored loads and the resistance factors alone. A total
comparison would have to also include any advances in more realistic
analysis methods and resistance formulations which are included in
the LRFD Specification. Some of these features have been, and more
may be, adopted into the Standard Specifications as time goes on.
printed on June 24, 2003
Lecture - 1-36
Figure 1.4.6.3-2 - Structural Demand - LRFD vs. Standard
Specification
printed on June 24, 2003
Lecture - 1-37
REFERENCES
American Association of State Highway and Transportation Officials,
"Standard Specifications for Highway Bridges", Thirteenth Edition,
1983, as amended by the Interim Specifications - Bridges, 1984, 1985,
Washington, D. C.
British Standards Institute, "British Standard 5400 - Parts 1 through
10", London (1978-1984)
Dorton, R. A., "Implementing the New Ontario Bridge Code",
International Conference on Short- and Medium-Span Bridges,
Toronto (1982)
Gellert, W., Kustner, H., Hellwich, M., and Kastner, H. - Editors, The
VNR Concise Encyclopedia of Mathematics, Van Nostrand Reinhold
(1977) pp. 591-600
Harr, M. E., Reliability-Based Design in Civil Engineering, McGraw-Hill
Book Company (1987) pp. 3-4
International Standards Organization, ISO 2394, "General Principles
for the Verification of the Safety of Structures"
Japan Road Association "Specifications for Highway Bridges, Part I:
Common Specifications and Part III: Concrete Bridges" (March 1984)
Nowak, A. S. and Lind, N. C., "Practical Bridge Code Calibration",
ASCE, Journal of the Structural Division, Vol. 105, No. ST12
(December 1979) pp. 2497-2510
Nowak, A. S., "Calibration of LRFD Bridge Design Code", Department
of Civil and Environmental Engineering Report UMCE 92-25,
University of Michigan, 1993
Ontario Ministry of Transportation and Communications, "Ontario
Highway Bridge Design Code", Toronto, Ontario, Canada (1983)
Vincent, G. S., "Load Factor Design of Steel Highway Bridges", AISI
Bulletin 15 (March 1969)
printed on June 24, 2003
Lecture - 2-1
LECTURE 2 - GENERAL FEATURES OF DESIGNS
2.1 OBJECTIVE OF THE LESSON
The objective of this lesson is to familiarize the student with the
contents of the general features portion of the LRFD Specification.
This section of the Specification sets forth a series of safety,
geometric, hydraulic and other requirements that need to be
considered when establishing the site for a bridge.
Further, it defines the objectives of good design in very general
terms. The intent of this section was more than simply a matter of
establishing certain "laundry lists" to be checked off during design.
Rather, it was a reaction to the compartmentization of the design
process which sometimes separates the Bridge Engineer, the
Hydraulics Engineer, the Geotechnical Engineer, the Transportation
Planner and the Maintenance Engineer from each other, rather than
allowing them to work collegially to establish the best overall solution
to a transportation system need. It was hoped that reminding the
Bridge Engineer of these collegial requirements in the bridge
specification would enable the Bridge Engineer to be injected earlier
into the overall consideration of planning route design. If all
components of the design team work together, then not only will the
important hydraulic, geometric and other requirements be met, but
they can be met also accounting for and accommodating, where
possible, simplified bridge geometrics.
2.2 LOCATION FEATURES
This section of the Specification provides general guidance on
bridge siting and introduces the general requirements for crossing
flood plains and other waterways. Bridges and bridge approaches on
flood plains are required to be compatible with the goals and
objectives of flood plain management as practiced by the various
states. Basically, it is a question of identifying and mitigating the
hydraulic consequences of having substructures units and roadway
embankments in areas apt to be flooded by various design floods.
Similarly, the effect of engineered construction on the long-term
aggradation or degradation of channel boundaries and depths must be
considered, both on a local, i.e., local scour, level, but on a water
course level as well.
The safety considerations related to the bridge as an
obstruction to land and marine traffic must also be considered.
Various guidelines in the AASHTO Roadside Design Guide are
available to assist the Bridge Designer in determining when
substructure units must be protected by guardrail or other barriers.
Provisions for confining or redirecting errant vehicles on the roadway,
supported by the bridge, are provided in Section 13 of the
Specifications which deals with barriers and railings.
printed on June 24, 2003
Lecture - 2-2
Wherever possible and practical, bridges crossing navigable
waterways should be designed with piers outside of the waterway.
Where the economic consequences of doing this are severe, then
bridges must be protected against vessel collision by suitable fenders,
dikes or dolphins. Guidance on the design of dolphin and fender
systems, as well as ship collision forces, are given in Section 3 of the
Specification. Further guidance can be obtained from several
AASHTO publications, including highway drainage guidelines,
hydraulic analysis for the location of design of bridges and the
AASHTO Guide Specification for vessel passage.
Vertical and horizontal clearances for highway bridges, and
clearances for railroad overpasses are identified in the Specification.
Generally speaking, these clearances will be consistent with the
AASHTO policy on geometric design of highway and streets and the
AREA Manual for Railway Engineering, Chapters, 7, 8, 9, 15 and 18.
Various numerical data are presented in this portion of the
Specification, which may be subject to jurisdictional differences, but
the following general principles should apply:
Reductions and vertical clearances, due to settlement,
overlaying of roadways being crossed or other factors should
be included in establishing the overall clearance requirement.
Somewhat higher clearances should be established for signs
attached to structures and for the vertical clearance to
overhead structural members, such as truss portals.
No object on or under a bridge other than barriers should be
located closer than 1200 mm to the edge of the designated
lane, and the inside face of the barrier should not be closer
than 600 mm to either the face of the object or the edge of the
designated traffic light.
2.3 FOUNDATION INVESTIGATIONS
Foundation investigations are a necessary part of the siting of
any structure. Article 10.4 of the Specification provides a list of
possible requirements of a subsurface investigation program.
2.4 DESIGN OBJECTIVES
This group of articles in the Specification identifies a series of
objectives for all bridge design. They are: safety, serviceability,
constructibility, economy and bridge aesthetics.
printed on June 24, 2003
Lecture - 2-3
2.4.1 Safety
The Specification clearly establishes that public safety is the
primary responsibility of the Design Engineer. All other aspects of
design, including serviceability, maintainability, economics and
aesthetics are secondary to the requirement for safety. This does not
mean that other objectives are not important, but safety is paramount.
All comprehensive design specifications are written to establish
an acceptable level of safety, as explained in Lecture 1. There are
many methods of attempting to quantify safety and the method
inherent in the LRFD Specification is probability-based reliability
analysis. The design vehicle for treating safety issues in the LRFD
Specification is the establishment of "limit states" to define groups of
events or circumstances which could cause a structure to be
unserviceable for its original intent.
2.4.1.1 LIMIT STATES
The LRFD Specification is written in a probability-based limit
state format requiring examination of some, or all, of the four limit
states defined below for each design component of a bridge.
Service Limit States
Service limit states are restrictions on stress, deformation and
crack width under regular service conditions. They are intended to
allow the bridge to perform acceptably for its service life.
Fatigue and Fracture Limit States
Fatigue and fracture limit states are restrictions on stress range
under regular service conditions reflecting the number of expected
stress range excursions. They are intended to limit crack growth
under repetitive loads to prevent fracture during the design life of the
bridge.
Strength Limit States
Strength limit states are intended to ensure that strength and
stability, both local and global, are provided to resist the statistically
significant load combinations that a bridge will experience in its design
life. Extensive distress and structural damage may occur under
strength limit states, but overall structural integrity is expected to be
maintained.
Extreme Event Limit States
Extreme event limit states are intended to ensure the structural
survival of a bridge during a major earthquake, or when collided by a
vessel, vehicle or ice flow, or where the foundation is subject to the
scour which would accompany a flood of extreme recurrence, usually
considered to be 500 years. They are considered to be unique
printed on June 24, 2003
Lecture - 2-4
occurrences whose return period is significantly greater than the
design life of the bridge.
Limit States Design Equation
As the basis of LRFD methodology, each component and
connection are to satisfy Equation 2.4.1.1-1 for each limit state. All
limit states are of equal importance.

i

i
Q
i
R
n
= R
r
(2.4.1.1-1)
where:

i
=
D

R

I
: =
D

R

I
0.95 for loads for which a
maximum value of
i
is appropriate, and,
for loads for which a minimum

i
1

R
1.0
value of
i
is appropriate

i
= Load Factor: A statistically based multiplier on force
effects
= Resistance Factor: A statistically based multiplier
applied to nominal resistance

i
= Load Modifier

D
= A factor relating to ductility

R
= A factor relating to redundancy

I
= A factor relating to operational importance
Q
i
= Nominal Force Effect: A deformation, stress or stress
resultant
R
n
= Nominal Resistance: Based on the dimensions as
shown on the plans and on permissible stresses,
deformations or specified strength of materials
R
r
= Factored Resistance: R
n
Ductility, redundancy and operational importance are
significant aspects affecting the margin of safety of bridges. While the
first two directly relate to the physical strength, the last concerns the
consequences of the bridge being out of service. The grouping of
these aspects is, therefore, arbitrary, however, it constitutes a first
effort of codification. In the absence of more precise information, each
effect, except that for fatigue and fracture, is estimated as 5%,
accumulated geometrically, a clearly subjective approach. With time,
improved quantification of ductility, redundancy and operational
importance, and their interaction, may be attained.
Ductility
The response of structural components or connections beyond
the elastic limit can be characterized by either brittle or ductile
behavior. Brittle behavior is undesirable because it implies the sudden
loss of load carrying capacity immediately when the elastic limit is
exceeded. Ductile behavior is characterized by significant inelastic
printed on June 24, 2003
Lecture - 2-5
deformations before any loss of load carrying capacity occurs. Ductile
behavior provides warning of structural failure by large inelastic
deformations. Under cyclic loading, large reversed cycles of inelastic
deformation dissipate energy and have a beneficial effect on structure
response.
If, by means of confinement or other measures, a structural
component or connection made of brittle materials can sustain
inelastic deformations without significant loss of load carrying capacity,
this component can be considered ductile. Such ductile performance
shall be verified by experimental testing.
In order to achieve adequate inelastic behavior, the system
should have a sufficient number of ductile members and either:
joints and connections which are also ductile and can provide
energy dissipation without loss of capacity; or,
joints and connections which have sufficient excess strength
so as to assure that the inelastic response occurs at the
locations designed to provide ductile, energy absorbing
response.
Behavior which is ductile in a static context, but which is not
ductile during dynamic response, should also be avoided. Examples
of this behavior are shear and bond failures in concrete members, and
loss of composite action in flexural members.
Past experience generally indicates that typical bridge
components designed in accordance with these specifications exhibit
adequate ductility. Connection and joints require special attention to
detailing and the provision of adequate load paths.
For unusual and important structures in high-seismic zones,
the Owner may specify a minimum ductility factor as a demonstration
that ductile failure modes will be obtained. The factor may be defined
as:
(2.4.1.1-2)

y
where:

u
- Deformation at ultimate

y
- Deformation at the elastic limit
The ductility capacity of structural components or connections
may either be established by full or large scale experimental testing,
or with analytical models which are based on realistic material
behavior. The ductility capacity for a structural system may be
printed on June 24, 2003
Lecture - 2-6
determined by integrating local deformations over the entire structural
system.
The special requirements for energy dissipating devices are
imposed because of the rigorous demands placed on these
components.
Given proper controls on the innate ductility of basic materials,
proper proportioning and detailing of a structural system are the keys
consideration in ensuring the development of significant, visible,
inelastic deformations, prior to failure, at the strength and extreme
event limit states.
It may be assumed that the requirements for ductility are
satisfied in a concrete structure in which the resistance of a
connection is not less than 1.3 times the maximum force imposed on
the connection by the inelastic and ductile actions of the adjacent
components.
For the fatigue and fracture limit state for fracture-critical
members and for the strength limit state for all members:

D
1.05 for non-ductile components and connections,
= 1.00 for conventional designs and details complying
with these specifications
0.95 for components and connections for which
additional ductility-enhancing measures have been
specified beyond those required by these
Specifications
For all other limit states:

D
= 1.00
Redundancy
Multiple load path structures should be used, unless there are
compelling reasons to the contrary.
For the strength limit state:

R
1.05 for non-redundant members
= 1.00 for conventional levels of redundancy
0.95 for exceptional levels of redundancy
For all other limit states:

R
= 1.00
printed on June 24, 2003
Lecture - 2-7
Operational Importance
The concept of operational importance is applied to the
strength and extreme-event limit states. The Owner may declare a
bridge or any structural component and connection, thereof, to be of
operational importance. If a bridge is deemed of operational
importance,
I
is taken as 1.05. Otherwise,
I
is taken as 1.0 for
typical bridges and may be reduced to 0.95 for relatively less
important bridges.
Such classification should be based on social/survival and/or
security/defense requirements.
2.4.1.2 LOAD FACTORS AND LOAD COMBINATIONS
The following permanent and transient loads and forces shall
be considered:
Permanent Loads
DD = Downdrag
DC = Dead load of structural components
attachments
DW = Dead load of wearing surfaces and utilities
EF = Dead load of earth fill
EH = Horizontal earth pressure
ES = Earth surcharge load
EV = Vertical earth pressure
Transient Loads
BR = Vehicular braking force
CE = Vehicular centrifugal force
CR = Creep
CT = Vehicular collision force
CV = Vessel collision force
EQ = Earthquake
FR = Friction
IC = Ice load
IM = Vehicular dynamic load allowance
LL = Vehicular live load
LS = Live load surcharge
PL = Pedestrian live load
SE = Settlement
SH = Shrinkage
TG = Temperature gradient
TU = Uniform temperature
WA = Water load and stream pressure
WL = Wind on live load
WS = Wind load on structure
The load factors for various loads, comprising a design load
combination, are indicated in Table 2.4.1.2-1. The larger of the two
values for load factors shown for TU, TG, CR, SH and SE are to be
used when calculating deformations; the smaller value shall be used
printed on June 24, 2003
Lecture - 2-8
when calculating all other force effects. For each load combination,
every load that is indicated, including all significant effects due to
distortion, should be multiplied by the appropriate load factor. All
relevant subsets of the load combinations in Table 2.4.1.2-1 should be
investigated.
The factors should be selected to produce the total extreme
factored force effect. For each load combination, both positive and
negative extremes should be investigated. In load combinations
where one force effect decreases the effect of another, the minimum
value should be applied to the load reducing the force effect.
For permanent force effects, the load factor which produces
the more critical combination shall be selected from Table 2.4.1.2-2.
In the application of permanent loads, force effects for each of the
specified six load types should be computed separately. Assuming
variation of one type of load by span, length or component within a
bridge is not necessary. For each force effect, both extreme
combinations may need to be investigated by applying either the high
or the low load factor as appropriate. The algebraic sums of these
products are the total force effects for which the bridge and its
components should be designed. This article reinforces the traditional
method of selecting load combinations to obtain realistic extreme
effects, and is intended to clarify the issue of the variability of
permanent loads and their effects.
When the permanent load increases the stability or load-
carrying capacity of a component or bridge, the minimum value of the
load factor for that permanent load shall also be investigated. Uplift,
which is treated as a separate load case in past editions of the
AASHTO Standard Specification for Highway Bridges, becomes a
Strength I load combination. For example, when the dead load
reaction is positive and live load can cause a negative reaction, the
load combination would be 0.9DC + 0.65DW + 1.75(LL+IM). If both
reactions were negative, the load combination would be 1.25DC +
1.50DW + 1.75(LL+IM).
The various load combinations shown in Table 2.4.1.2-1 are
described below.
STRENGTH I - Basic load combination relating to the normal
vehicular use of the bridge without wind.
STRENGTH II - Load combination relating to the use of the
bridge by permit vehicles without wind. If a
permit vehicle is traveling unescorted, or if
control is not provided by the escorts, the other
lanes may be assumed to be occupied by the
vehicular live load herein specified. For
bridges longer than the permit vehicle, addition
of the lane load, preceding and following the
permit load in its lane, should be considered.
printed on June 24, 2003
Lecture - 2-9
STRENGTH III - Load combination relating to the bridge
exposed to maximum wind velocity which
prevents the presence of significant live load
on the bridge.
STRENGTH IV - Load combination relating to very high dead
load to live load force effect ratios. This
calibration process had been carried out for a
large number of bridges with spans not
exceeding 60 m. Spot checks had also been
made on a few bridges up to 180 m spans. For
the primary components of large bridges, the
ratio of dead and live load force effects is
rather high, and could result in a set of
resistance factors different from those found
acceptable for small- and medium-span
bridges. It is believed to be more practical to
investigate one more load case, rather than
requiring the use of two sets of resistance
factors with the load factors provided in
Strength Load Combination I, depending on
other permanent loads present. This Load
Combination IV is expected to govern when the
dead load to live load force effect ratio exceeds
about 7.0.
STRENGTH V - Load combination relating to normal vehicular
use of the bridge with wind of 90 km per hour
velocity.
EXTREME EVENT - Load combination relating to ice load, collision
by vessels and vehicles, and to certain
hydraulic events and earthquakes whose
recurrence interval exceeds the design life.
The joint probability of these events is
extremely low, and, therefore, they are
specified to be applied separately. Under
these extreme conditions, the structure is
expected to undergo considerable inelastic
deformation by which locked-in force effects
due to TU, TG, CR, SH and SE will be relieved.
The 0.50 live load factor signifies a low
probability of the presence of maximum
vehicular live load at the time when extreme
events may occur.
SERVICE I - Load combination relating to the normal
operational use of the bridge with 90 km per
hour wind. All loads are taken at their nominal
values and extreme load conditions are
excluded.
printed on June 24, 2003
Lecture - 2-10
SERVICE II - Load combination whose objective is to prevent
yielding of steel structures due to vehicular live
load, approximately halfway between that used
for Service I and Strength I limit state, for which
case the effect of wind is of no significance.
This load combination corresponds to the
overload provision for steel structures in past
editions of the AASHTO Standard Specification
for the Design of Highway Bridges.
SERVICE III - Load combination relating only to prestressed
concrete structures with the primary objective
of crack control. The addition of this load
combination followed a series of trial designs
done by 14 states and several industry groups
during 1991 and early 1992. Trial designs for
prestressed concrete elements indicated
significantly more prestressing would be
needed to support the loads specified in the
proposed Specifications. There is no
nationwide physical evidence that these
vehicles used to develop the notional live loads
have caused detrimental cracking in existing
prestressed concrete components. The
statistical significance of the 0.80 factor on live
load is that the event is expected to occur
about once a year for bridges with two design
lanes, less often for bridges with more than two
design lanes, and about once a day for the
bridges with a single design lane.
FATIGUE - Fatigue and fracture load combination relating
to gravitational vehicular live load and dynamic
response, consequently BR and PL need not
be considered. The load factor reflects a load
level which has been found to be
representative of the truck population, with
respect to large number of return cycles.
Table 2.4.1.2-1 - Load Combinations and Load Factors
Load Combination
Limit State
DC
DD
DW
EH
EV
ES
LL
IM
CE
BR
PL
LS
WA WS WL FR TU
CR
SH
TG SE Use One of These at a
Time
EQ IC CT CV
STRENGTH-I
p
1.75 1.00 - - 1.00 0.50/1.20
TG

SE
- - - -
STRENGTH-II
p
1.35 1.00 - - 1.00 0.50/1.20
TG

SE
- - - -
STRENGTH-III
p
- 1.00 1.40 - 1.00 0.50/1.20
TG

SE
- - - -
printed on June 24, 2003
Lecture - 2-11
STRENGTH-IV
EH, EV, ES, DW
DC ONLY

p
1.5
- 1.00 - - 1.00 0.50/1.20
- -
- - - -
STRENGTH-V
p
1.35 1.00 0.40 0.40 1.00 0.50/1.20
TG

SE
- - - -
EXTREME EVENT-I
p

EQ
1.00 - - 1.00 - - - 1.00 - - -
EXTREME EVENT-II
p
0.50 1.00 - - 1.00 - - - - 1.00 1.00 1.00
SERVICE-I 1.00 1.00 1.00 0.30 0.30 1.00 1.00/1.20
TG

SE
- - - -
SERVICE-II 1.00 1.30 1.00 - - 1.00 1.00/1.20 - - - - - -
SERVICE-III 1.00 0.80 1.00 - - 1.00 1.00/1.20
TG

SE
- - - -
FATIGUE-LL, IM &
CE ONLY - 0.75 - - - - - - - - - - -
Table 2.4.1.2-2 - Load Factors for Permanent Loads,
p
Load Factor
Type of Load Maximum Minimum
DC: Component and Attachments 1.25 0.90
DD: Downdrag 1.80 0.45
DW: Wearing Surfaces and
Utilities
1.50 0.65
EH: Horizontal Earth Pressure
Active
At-Rest
1.50
1.35
0.90
0.90
EV: Vertical Earth Pressure
Overall Stability
Retaining Structure
Rigid Buried Structure
Rigid Frames
Flexible Buried Structures
other than Metal Box
Culverts
Flexible Metal Box Culverts
1.35
1.35
1.30
1.35
1.95
1.50
N/A
1.00
0.90
0.90
0.90
0.90
ES: Earth Surcharge 1.50 0.75
2.4.2 Serviceability
The Specification treats serviceability from the view points of
durability, inspectibility, maintainability, rideability, deformation control
and future widening.
The discussion of durability in Article 2.5.2 et sequa consists
basically of indicating that contract documents should call for high
quality materials and that those materials that are subject to
deterioration from moisture content and/or salt attack be protected.
printed on June 24, 2003
Lecture - 2-12
Inspectibility is to be assured through adequate means for permitting
inspectors to view all parts of the structure which have structural or
maintenance significance. The provisions related to inspectibility are
relatively short, but as all Departments of Transportation have begun
to realize, bridge inspection can be very expensive and is a recurring
cost due to the need for biennial inspections. Therefore, the cost of
providing walkways and other access means and adequate room for
people and inspection equipment to be moved about on the structure
is usually a good investment.
Maintainability is treated in the Specification in a similar
manner to durability; there is a list of desirable attributes to be
considered.
The subject of live load deflections and other deformations,
treated in Article 2.5.2.6, was a very difficult issue for the writers of the
Specification to come to grips with. On the one hand, there is very
little direct correlation between live load deflection and premature
deterioration of bridges. There is much speculation that "excessive"
live load deflection contributes to premature deck deterioration, but, to-
date (late 1993), no causative relationship has been statistically
established.
Rider comfort is often advanced as a basis for deflection
control. Many studies in human response to motion have shown that
it is not the magnitude of the motion, but rather the acceleration that
most people perceive, especially in moving vehicles. Many people
have experienced the sensation of being on a bridge and feeling a
definite movement, especially when traffic is stopped. Many times,
this movement is related to the movement of floor systems, which are
really quite small in magnitude, but noticeable nonetheless. There
being no direct correlation between magnitude (not acceleration) of
movement and discomfort has not prevented the design profession
from finding comfort in controlling the gross stiffness of bridges
through a deflection limit. In earlier drafts of the proposed LRFD
Specification, deflection criteria and other measures of gross stiffness
were placed in the commentary as historical information. During the
first trial design reporting meeting, a straw vote was taken of the state
bridge engineers in attendance as to whether some sort of gross
stiffness controls should be reinserted into the Specification, rather
than retaining it in the commentary. The bridge engineers in
attendance voted unanimously that it was their preference that some
sort of deflection criteria be reestablished. As a compromise between
the need for establishing comfort levels and the lack of compelling
evidence that deflection, per se, was cause of structural distress, the
deflection criteria were reestablished as voluntary provisions to be
activated by those states that so chose. Deflection limits, stated as
span divided by some number, were established for most cases, and
additional provisions of absolute relative displacement between planks
and panels of wooden decks and ribs of orthotropic decks were also
added. Similarly, optional criteria were established for a span-to-depth
ratio for guidance primarily in starting preliminary designs, but also as
printed on June 24, 2003
Lecture - 2-13
a mechanism for checking when a given design deviated significantly
from past successful practice.
2.4.3 Constructibility
Several new provisions were added in this article related to:
the need to design bridges so that they can be fabricated and
built without undue difficulty and with control over locked in
construction force effects,
the need to document one feasible method of construction in
the contract documents, unless the type of construction is
self-evident, and
a clear indication of the need to provide strengthening and/or
temporary bracing or support during erection, but not requiring
the complete design thereof.
2.4.4 Bridge Aesthetics
The treatment of bridge aesthetics is a difficult issue because
beauty is really in the eyes of the beholder. Reference is provided to
several guidelines and surveys on bridge aesthetics and a group of
starting concepts are given.
2.4.5 Hydrology and Hydraulics
Provisions in this article relate to the minimum considerations
of site-specific data collection, the extension to flow rates, stream
stability determining bridge waterway openings and the effects of
hydraulics on bridge foundations. Among the provisions are:
Temporary structures have to be designed for hydraulic, as
well as structural requirements. Contract documents for the
temporary structures have to delineate the respective
responsibilities and risks to be assumed by the highway
agency and the Contractor.
The specification of flood flows which should be investigated
for assessing flood hazards and meeting flood plain
requirements, assessing risks to users and damage to bridges
and roadway approaches, assessing catastrophic damage at
high risk sites, investigating adequacy of foundations to resist
scour and other factors.
Considerations with respect to stream stability and bridge
waterway opening are also delineated.
Two floods are identified for bridge scour. The first is the
design flood for scour based on the more severe of the 100-
year flood or an overtopping flood, in which case the structure
must be shown to be adequate for all applicable strength and
printed on June 24, 2003
Lecture - 2-14
service limit state load combinations identified in Table 2.4.1.2-
1 of the Specification. The second check flood for scour is
defined as a flood not to exceed a 500-year event or the
overtopping flood, in which case the structure need only meet
the extreme event limit state load combinations in the same
table. The indicated floods are a minimum requirement for
investigation, and the Engineer may also use a more severe
flood if local site conditions seem to justify that in the analysis
module.
printed on June 24, 2003
Lecture - 3-1
LECTURE 3 - LOADS - I
3.1 OBJECTIVE OF THE LESSON
The objective of this lesson is to acquaint the student with:
the concept of limit states as it applies to the application of the
LRFD Specification;
the expression of the limit states through the use of load
combinations;
the development of the LRFD live load model through the dual
approach of modeling the force effects generated by a group
of vehicles permitted on the roads and highways of various
states without special permits or other controls, herein referred
to as exclusion loads, and the use of statistical procedures to
project a 75-year occurrence live load;
the live load model in the U. S. LRFD Specification, and a
comparison to that contained in the Ontario Highway Bridge
Design Code, Third Edition;
An example of the application of the LRFD load model follows.
3.2 DEVELOPMENT OF LRFD LIVE LOAD MODEL
3.2.1 Background
Early editions of the AASHTO Standard Specification for
Highway Design contained a representation of a truck and/or a group
of trucks for use in design. The very earliest editions of this
Specification contained a single unit truck weighing up to 178 kN (20
U. S. customary tons), which was known as the H20 truck. Lighter
variations of this vehicle were also considered and were designated
as HXX, e.g., H15. Groups of 133 kN trucks, with an occasional 178
kN truck, were also utilized as a truck-train.
In 1931, the first edition of the Standard Specification instituted
the "lane load", which consisted of a uniform load of 9.3 N per mm and
a moving concentrated load or loads. A concentrated load of 116 kN
was used for shear and for reaction, two 80 kN concentrated loads
were used for negative moment at a support and were positioned in
two adjacent spans, and a single 80 kN load was used for all other
moment calculations.
In 1944, the truck was extended into a tractor-semi-trailer
combination, known as the HS20-44, commonly referred to as simply
the HS20 truck. This vehicle weighed a total of 325 kN and was
comprised of a single steering axle weighing 35 kN and two axles that
supported the semi-trailer, each weighing 145 kN. The axles on the
printed on June 24, 2003
Lecture - 3-2
semi-trailer could vary from 4.3 m to 9.0 m, and it was assumed that
there was 4.3 m between the steering axle and the adjacent axle that
formed part of the tractor. These loads are shown in Figure 3.2.1-1.
The HS20 truck was an idealization and did not represent one
particular truck, although it was clearly indicative of the group of
vehicles commonly known as 3-S2's, e.g., the common "18-wheeler".
Figure 3.2.1-1 - HS20 Truck Loading
During the late 1970s and early 1980s, some states raised
their design load to HS25, or 125% of some or all of the loads shown
in Figure 3.2.1-1. During the construction of the interstate system, an
additional design load, known as the "Interstate Load", was also
introduced, and this consisted of two 110 kN axles separated by 1.2
m. Some of the states which raised the design load to HS25 also
increased the weight of the design tandem to 133 kN.
The actual configurations of trucks are almost limitless, and it
has long been recognized that the regulatory control of vehicles is a
significantly different matter than the choice of a design model upon
which to base calculations. The link between these two needs are
state legal loads and the bridge formula. These two regulatory
devices have the objective of recognizing that the commercial needs
of the Country can only be satisfied by a plethora of vehicle
configurations and that it is sometimes necessary that these
configurations create significantly larger force effects in structures than
those calculated using the design model. The AASHTO Manual for
Maintenance Inspection of Bridges has long recognized this by
prescribing two stress levels commonly referred to as the Inventory
Level, which approximates design stresses or design force effects and
the Operating Level, which results in approximately 1/3 more total
force effect on the basis that it will occur, although relatively
infrequently. The range between these two stress levels is often used
as a basis of establishing fleets of vehicles which produce more force
effect than legal loads, but for which permits will be routinely granted
without the need for individual structural calculations.
Over the years, many states have written exclusions into their
regulatory policies, which permitted some vehicles in excess of legal
loads to operate in an unrestricted manner. These loads are
sometimes referred to as "grandfather provision" loads.
As a result of the evolution summarized above, it became
increasingly apparent that the HS loading did not bear a uniform
relationship to many of the vehicles allowed on the roads. It was
becoming out-of-date. In developing a new design specification, it
became apparent quite early in the development process that if the
printed on June 24, 2003
Lecture - 3-3
objective of developing a new specification was a more uniform and
consistent safety of bridges, a new live load model would be
necessary in order to produce that consistency.
3.2.2 Selection of a Basis for Developing a Model
As outlined in Section 3.2.1, the current regulatory situation
embodied in state legal loads, unanalyzed permit loads, grandfather
provisions and the bridge formula, could provide one basis for
identifying live load force effects, which could be extended into a
design load. In 1990, the Transportation Research Board published
Special Report 225 entitled, "Truck Weight Limits - Issues and
Options", summarizing a study into the state legal loads, grandfather
provisions, the current bridge formula and various attempts to extend
the bridge formula, or to develop other regulatory models. This study
contained extensive information on the estimated benefits from more
efficient movement of goods compared to the cost and accelerated
damage to roadways and bridges. A group of vehicles were identified
in Special Report 225 and were made available to the LRFD
development group as part of personal correspondence. Various
vehicle configurations arising from this work are shown in Figure
3.2.2-1.
A proposal by the National Truck Weight Action Committee
(NTWAC) was reduced to three special hauling vehicles
(SHVs) shown in Figure 3.2.2-1. These vehicles can be
represented as a three-axle single unit weighing 356 kN, a
four-axle single unit with a tridem axle unit weighing a total of
378 kN and a 3-S-2 weighing 489 kN.
Vehicles representative of a project to produce a modified
bridge formula conducted by the Texas Transportation Institute
(TTI), can be embodied in the four configurations shown in
Figure 3.2.2-1, which are similar in shape, but not weight, to
the NTWAC vehicles and are extended by one additional 421
kN 3-S2-2.
The Canadian Interprovincial Loads, which resulted from a
1988 agreement by the Canadian Council of Ministries of
Transportation and Highway Safety, produced a common set
of weight limits (RTAC 1988) for tractor-semi-trailers and
double trailer combinations, as shown in Figure 3.2.2-1. Some
of these vehicles are similar to the TTI vehicles. An additional
axle is added to the 3-S2-2, and the spacings and weights of
the axles are somewhat different, and a fifth configuration,
called a 3-S2-4, is also added.
The Extended Bridge Formula vehicles, shown in Figure 3.2.2-
1, are long combination vehicles (LCVs) intended to extend the
bridge formula past 356 kN.
Turner trucks, two vehicle combinations known as the Turner
A and Turner B Trucks, shown in Figure 3.2.2-1, which were
printed on June 24, 2003
Lecture - 3-4
developed on the principle that pavement would be less
damaged by vehicles with increased gross vehicle weight if the
weight per axle was reduced by adding additional axles.
The vehicles shown in Figure 3.2.2-1 under the heading
AASHTO Rating Vehicles were thought to be "typical legal
load types" used by many states for rating, instead of the
HS20 load configuration.
Vehicle configurations representing the grandfather provision
exclusions to legal loads available in various states are shown
as "EX" vehicles in Figure 3.2.2-1. They represent various
types of special hauling vehicles and long combination
vehicles common in the United States. This "family" of
vehicles is referred to herein as "Exclusion Loads".
printed on June 24, 2003
Lecture - 3-5
Figure 3.2.2-1 - Truck Configuration Considered
printed on June 24, 2003
Lecture - 3-6
Figure 3.2.2-1 - Truck Configuration Considered (Continued)
printed on June 24, 2003
Lecture - 3-7
The force effect on bridge structures from these various
"families" of vehicles were studied by calculating the envelope of force
effects from each of the representative vehicles in a "family" for:
Centerline moment of a simply-supported beam
Positive and negative moment at the 0.4L point of a two-span
continuous girder, with two equal spans
End shear and shear at both sides of the interior support of a
two-span continuous girder, with two equal spans
Figure 3.2.2-2 shows a comparison of the various moment-type force
effects, identified above, for spans from 6 to 45 m generated by the
exclusion vehicles compared to the HS20 truck. This comparison is
developed by plotting the ratio of the force effect from the envelope of
exclusion vehicles divided by the corresponding force effect from the
HS20 vehicle on a vertical axis, against span length on the horizontal
axis. Thus, a complete match of force effects, indicating that the
HS20 vehicle was an accurate and representative model of the
exclusion loads, would be indicated by a horizontal line passing
through the vertical axis at a value of 1.0. Corresponding information
for the shear force effects identified above are shown in Figure 3.2.2-
3, in which V
ab
is the shear at the simply-supported end and V
ba
is the
shear adjacent to the interior support. Figures 3.2.2-2 and 3.2.2-3,
taken together, show that the HS20 vehicle is not representative of
current loads on the highways and documents the need to develop a
new live load model.
Figure 3.2.2-2 - EXCL/HS20 Truck or Lane or 2 - 110 kN Axles @ 1.2
m
printed on June 24, 2003
Lecture - 3-8
Figure 3.2.2-3 - EXCL/HS20 Truck or Lane or 2 - 110 kN Axles @ 1.2
m
The following notation applies to Figures 3.2.2-4 through 3.2.2-
10:
Figures 3.2.2-4 through 3.2.2-10 compare the force effects
created by the HS20 truck, the Turner trucks, the TTI trucks, the
extended bridge formula trucks and the exclusion trucks for various
span lengths and expressed in either kN or kN m, as appropriate. The
AASHTO rating vehicles were not found to govern these conditions
and have not be plotted on the figures.
In the case of negative moment of an interior support, the
results shown in these figures correspond to one vehicle on the span.
Additional studies evaluated the effects of two vehicles on the span.
Note that in the case of negative moment support, the HS loading is
often actually controlled by the lane load and a pair of concentrated
loads, and hence, it is more representative of that force effect than the
HS loading appears to be for moment at the centerline of a simple
span, as shown in Figure 3.2.2-4.
printed on June 24, 2003
Lecture - 3-9
Figure 3.2.2-4 - Centerline Moments - Simple Span
Figure 3.2.2-5 - Negative Moments at 0.4L
Figure 3.2.2-6 - Negative Moments at Support
printed on June 24, 2003
Lecture - 3-10
Figure 3.2.2-7 - Positive Moments at 0.4L
Figure 3.2.2-8 - Positive Shear at + V-AB
Figure 3.2.2-9 - Negative Shear at -V-AB
printed on June 24, 2003
Lecture - 3-11
Figure 3.2.2-10 - Negative Shear at -V-BA
3.2.3 Candidate Notional Loads
Five candidate notional loads were identified early in the live
load development for the AASHTO LRFD Bridge Specification:
A single vehicle, called the HTL57, weighing a total of 507 kN
and having a fixed wheel base and a fixed axle spacing and
weights shown in Figure 3.2.3-1. This vehicle is similar to the
design vehicle contained in the 1983 Edition of the Ontario
Highway Bridge Design Code.
Figure 3.2.3-1 - HTL-57 (507 kN)
A "family" of three loads shown in Figure 3.2.3-2, consisting of
a tandem, a four-axle single unit, with a tridem rear
combination, and a 3-S-2 axle configuration taken together
with a uniform load, preceding and following that axle
grouping.
Figure 3.2.3-2 - Family of Three Loads
printed on June 24, 2003
Lecture - 3-12
A design "family" called "HL93" consisting of subsets or
combinations of a design tandem similar to that shown in
Figure 3.2.3-2, the HS20 truck shown in Figure 3.2.1-1, and a
uniform load of 9.3 N per running mm of lane, as shown in
Figure 3.2.3-3. Also shown in Figure 3.2.3-3 is an extension
of this loading to include 90% of two HS20 trucks and 90% of
the uniform load for the cases indicated.
Figure 3.2.3-3 - HL93 Design Load
Not shown is a slight variation of the combination of the HS
vehicle and the uniform load, which involves an HS25 load,
followed and preceded by a uniform load of 7.0 N per running
mm of lane, with the uniformly distributed load broken for the
HS vehicle.
An equivalent uniform load in N per mm of lane required to
produce the same force effect as the envelope of the exclusion
vehicles for various span lengths as shown in Figure 3.2.3-4.
printed on June 24, 2003
Lecture - 3-13
Figure 3.2.3-4 - Equivalent Uniform Load
Consideration of Figure 3.2.3-4 indicates that:
Each force effect would have to have its own equation for
uniform load
The equation of the uniform load would be nonlinear in span
length
In view of these complexities, an equivalent uniform load without
concentrated loads was eliminated.
A comparison of the remaining possible load configurations is
shown in Figures 3.2.3-5 through 3.2.3-11 for each of the considered
force effects.
The following notation applies to Figures 3.2.3-5 through 3.2.3-
11:
Data is presented as a ratio of a given force effect from the envelope
of exclusion loads divided by the corresponding force effect from each
of the live load models considered. Thus, a value of greater than 1.0
on the vertical axis indicates a situation in which the envelope of the
exclusion vehicles produce more force effect than the design model
under consideration. Ratios for the HS20 vehicle are also included for
reference. Once again, a horizontal line, intersecting the vertical axis
at a value of 1.0, indicates a perfect match. Generally speaking,
Figures 3.2.3-5 through 3.2.3-11 indicate that the load model involving
a combination of either a pair of 110 kN tandem axles and the uniform
load, or the HS20 and the uniform load, seem to produce the best fit
to the exclusion vehicles. This is summarized in tabular form in Table
3.2.3-1, in which the mean and standard deviations of the data shown
printed on June 24, 2003
Lecture - 3-14
in Figures 3.2.3-5 through 3.2.3-11 are given for each of the force
effects indicated for each of the models under consideration. This
summary also indicates that the model consisting of either the tandem
plus the uniform load, or the HS20 plus the uniform load, produce the
best results.
Table 3.2.3-1 - Live Load Models vs Exclusion Load Mean and
Standard Deviation
HS20
HS20+9.3
(Prop.) HS25+7.0 HTL FAMILY-3
HTL MOD
LF
-M 0.4L MEAN 1.600 1.060 0.982 1.189 1.048 1.091
STD
DEV
0.1679 0.0630 0.0755 0.1290 0.0379 0.0389
-M @SUPT MEAN 1.111 0.847 0.835 0.952 0.866 0.871
STD
DEV
0.2068 0.1201 0.0712 0.1427 0.0917 0.0575
+M 0.4L MEAN 1.459 1.018 0.924 1.200 1.041 1.103
STD
DEV
0.1186 0.0618 0.0409 0.1133 0.0475 0.0464
SIM SUP MEAN 1.506 1.001 0.941 1.198 1.050 1.100
STD
DEV
0.1387 0.0275 0.0586 0.1325 0.0363 0.0305
-V ab MEAN 1.544 1.060 0.982 1.189 1.048 1.092
STD
DEV
0.1253 0.0929 0.0755 0.1290 0.0379 0.0389
-V ba MEAN 1.461 1.011 0.932 1.111 1.006 1.022
STD
DEV
0.1448 0.0355 0.0475 0.1093 0.0522 0.0540
+V ab MEAN 1.415 1.024 0.919 1.132 1.017 1.232
STD
DEV
0.1447 0.0391 0.0391 0.1027 0.0558 0.2792
Figure 3.2.3-5 - Results for Simple Span Centerline Moments
printed on June 24, 2003
Lecture - 3-15
Figure 3.2.3-6 - Results for Negative Moments at 0.4L
Figure 3.2.3-7 - Results for Negative Moments at Support
Figure 3.2.3-8 - Results for Positive Moments at 0.4L
printed on June 24, 2003
Lecture - 3-16
Figure 3.2.3-9 - Results for Positive Shear at +V-AB
Figure 3.2.3-10 - Results for Negative Shear at -V-AB
Figure 3.2.3-11 - Results for Negative Shear at -V-BA
A summary of the force effect ratios for the exclusion vehicles
divided by either the tandem plus the uniform load, or the HS20 truck
printed on June 24, 2003
Lecture - 3-17
plus the uniform load, is shown in Figures 3.2.3-12 and 3.2.3-13. It
can be seen that the results for force effects under consideration are
tightly clustered, very parallel, and form bands of data which are
essentially horizontal. The tight clustering of data for the various force
effects indicates that one notional model can be developed for all of
the force effects under consideration. The fact that the data is
essentially horizontal indicates that both the model and the load factor
applied to live can be independent of span length. The tight clustering
of all the data for all force effects further indicates that one live load
factor will suffice.
Figure 3.2.3-12 - EXCL/HS20+9.3 or Dual 110 kN Moment Ratio
Figure 3.2.3-13 - EXCL/HS20+9.3 or Dual 110 kN Shear Ratio
Thus, the combination of the tandem with the uniform load and
the HS20 with the uniform load, were shown to be an adequate basis
for a notional design load in the LRFD Specification.
In mid-1993, the Third Draft of the OHBDC became available.
It is instructive to compare the force effects generated by the live load
model in that specification and the LRFD Specification. For this
comparison, the force effect are multiplied by the load factors required
printed on June 24, 2003
Lecture - 3-18
by the two specifications. This produces a more meaningful
comparison.
Figure 3.2.3-14 shows a comparison of the moments
presented as a ratio of LRFD divided by OHBDC, produced for various
span lengths for a simply-supported beam and a two-span continuous
unit of equal spans for the force effects indicated in the figure. Neither
dynamic load allowance or the multiple presence factor (MPF) are
taken into account in this figure, which shows that there is a relatively
high correlation between the force effects between the two load
models, even though they were developed by different working
groups. There is some difference in the very short spans, but the
major difference indicated is for negative moment at, or near, a
support. In this latter case, the LRFD Specification requires the use
of two design trucks. The OHBDC has no comparable provision.
Similar results for shear are shown in Figure 3.2.3-15, which also
show a high correlation between the shear force effects produced by
the two load models.
Of course, the total live load force effect is also a function of
the impact and multiple presence factors. Both of these factors are
different in the two specifications. Figure 3.2.3-16 shows moment
ratios for various spans for the case where both the dynamic load
allowance and the one-lane multiple presence factor are incorporated.
In this case, it will be seen that the LRFD Specification produces
considerably more force effect due to its multiple presence factor of
1.20 when only one lane is loaded. Comparable results for shear are
shown in Figure 3.2.3-17. Figures 3.2.3-18 through 3.2.3-23 present
additional studies of moment ratio and shear ratio for two lanes
loaded, three lanes loaded and four lanes loaded. It can be seen that
until four loaded lanes controls a design, the LRFD Specification
generally produces somewhat more force effect than the OHBDC.
When the case involving four-lane loaded lanes controls the design,
then the OHBDC will provide additional shear and moment, on the
order of 10%. One might assume that a bridge wide enough to have
four lanes control the design might very well be considered
"important", in which case the importance factor, in the AASHTO
specification,
i
, will be taken greater than one, which may offset this
difference, where appropriate.
The full comparison of specification really requires that both
the load factors and all other factors on live load, the dead load and
other appropriate loads be compared to resistance and resistance
factors. Generally speaking, the Ontario Specification has resistance
factors which are approximately 5% less than the LRFD Specification,
which will make the force effects generated by the two specifications
for one loaded lane, two loaded lanes and three loaded lanes, where
they govern, closer to each other.
printed on June 24, 2003
Lecture - 3-19
Figure 3.2.3-14 - HL-93/OHBD Moment Ratio
(no impact or MPF)
Figure 3.2.3-15 - HL-93/OHBD Shear Ratio (no impact or MPF)
Figure 3.2.3-16 - HL-93/OHBD Moment Ratio (with impact and one-
lane MPF)
printed on June 24, 2003
Lecture - 3-20
Figure 3.2.3-17 - HL-93/OHBD Shear Ratio (with impact and one-lane
MPF)
Figure 3.2.3-18 - HL-93/OHBD Moment Ratio (with impact and two-
lane MPF)
Figure 3.2.3-19 - HL-93/OHBD Shear Ratio (with impact and two-lane
MPF)
printed on June 24, 2003
Lecture - 3-21
Figure 3.2.3-20 - HL-93/OHBD Moment Ratio (with impact and three-
lane MPF)
Figure 3.2.3-21 - HL-93/OHBD Shear Ratio (with impact and three-
lane MPF)
Figure 3.2.3-22 - HL-93/OHBD Moment Ratio (with impact and four-
lane MPF)
printed on June 24, 2003
Lecture - 3-22
Figure 3.2.3-23 - HL-93/OHBD Shear Ratio (with impact and four-lane
MPF)
3.2.4 Statistical Basis of Live Load Model
3.2.4.1 INTRODUCTION
Traditionally, the static and dynamic effects of live load are
considered separately; therefore, this article covers only the static
component. Dynamic effects are treated in Section 3.3.7.
The effect of live load depends on many parameters, including
the span length, truck weight, axle loads, axle configuration,
transverse and longitudinal position of the vehicle on the bridge,
number of vehicles on the bridge (multiple presence), girder spacing,
and stiffness of structural components. Because of the complexity of
the model, the variation in load and load distribution properties are
considered separately.
The statical live load model was based on available truck
survey data. The data considered included weigh-in-motion (WIM)
measurements performed as part of the FHWA project (Goble, et al,
1991), weigh-in-motion measurements carried out as a part of
Michigan DOT project (Nowak and Nassif, 1991) and truck
measurements performed by the Ontario Ministry of Transportation
(Agarwal and Wolkowicz, 1976). Other available WIM data was
analyzed as part of NCHRP Project 12-28(11), Development of Site-
Specific Load Models for Bridge Rating. However, it was found that
the data collected in the mid 1980's by various states was not reliable,
with errors estimated at 30-40%. Therefore, the data base consisted
of the results of a truck survey performed in 1975 by the Ontario
Ministry of Transportation. That study covered about 10,000 selected
trucks, i.e., only trucks which appeared to be heavily loaded were
measured and included in the data base.
The uncertainties involved in the analysis are due to limitations
and biases in the survey data. Even though 10,000 trucks is a large
number, it is very small compared to the actual number of heavy
printed on June 24, 2003
Lecture - 3-23
vehicles in a 75-year life time. It is also reasonable to expect that
some extremely heavy trucks purposefully avoided the weighing
stations. A considerable degree of uncertainty is caused by the
unpredictability of future trends with regard to configuration of axles
and weights.
3.2.4.2 TRUCK SURVEY DATA
The study is based on the truck survey including 9,250 heavy
vehicles (Agarwal and Wolkowicz, 1976). The data includes truck
configuration represented by the number of axles and axle spacing,
axle loads and gross vehicle weight. For each truck in the survey,
bending moments, M, and shear forces, V, are calculated for a wide
range of spans. Simple spans and continuous two equal spans are
considered. Moments and shears are calculated in terms of the
Standard HS20 truck or lane loading, whichever governs (AASHTO
1989). The cumulative distribution functions (CDF) are plotted on
normal probability paper in Figure 3.2.4.2-1 for simple span moments,
Figure 3.2.4.2-2 for shears and Figure 3.2.4.2-3 for negative moments
at interior supports of two continuous spans for spans from 9 to 60 m.
The vertical scale, z, is:
(3.2.4.2-1) z
1
[F(x)]
where:
F(x) = cumulative distribution function of x, where x is the
moment M or shear
V
-1
= inverse of the standard normal distribution function
3.2.4.3 MEAN MAXIMUM TRUCK MOMENTS AND SHEARS
The mean maximum moments and shears for various time
periods are determined by extrapolation of the distributions, as shown
in Figures 3.2.4.3-1, 3.2.4.3-2 and 3.2.4.3-3 and as shown by the
heavy extensions to each curve. Let N be the total number of heavy
trucks in time period T. It is assumed that the surveyed trucks
represent about two-week traffic. Therefore, in T = 75 years, the
number of trucks, N, will be about 2,000 times larger than in the
survey. This will result in N = 20 million heavy trucks. The probability
level corresponding to N is 1/N, and for N = 20 million, it is
1/20,000,000 = 5
10
-8
, which corresponds to z = 5.33 on the vertical
scale, as shown in Figures 3.2.4.2-1, 3.2.4.2-2 and 3.2.4.2-3.
The number of trucks, N, probabilities, 1/N, and inverse normal
distribution values, z, corresponding to various time periods T from
one day to 75 years, are shown in Table 3.2.4.3-1.
The mean maximum moments and shears corresponding to
various periods of time can be read from the graph. For example, for
a 36 m span and T = 75 years, the mean maximum moment is equal
to 2.08 HS20 moment. The factor 2.08 is the horizontal coordinate of
printed on June 24, 2003
Lecture - 3-24
intersection of the extrapolated distribution and z = 5.33 on the vertical
scale.
Table 3.2.4.3-1 - Number of Trucks vs. Time
Period and Probability
Time
Period
T
Number of
Trucks
N
Probability
1/N
Inverse
Normal
z
75 years 20,000,000 5
10
-8
5.33
50 years 15,000,000 7
10
-8
5.27
5 years 1,500,000 7
10
-7
4.83
1 year 300,000 3
10
-6
4.50
6 months 150,000 7
10
-6
4.36
2 months 50,000 2
10
-5
4.11
1 month 30,000 3
10
-5
3.99
2 weeks 10,000 1
10
-4
3.71
1 day 1,000 1
10
-3
3.09
Figure 3.2.4.3-1 - Extrapolated Moments for Simple Spans
printed on June 24, 2003
Lecture - 3-25
Figure 3.2.4.3-2 - Extrapolated Shears for Simple Spans
Figure 3.2.4.3-3 - Extrapolated Negative Moments for Two Equal
Continuous Spans
printed on June 24, 2003
Lecture - 3-26
For comparison, the number of heavy trucks passing through the
bridge in five years is about 1,500,000. This corresponds to z = 4.83
on the vertical scale of Figure 3.2.4.3-1, and the resulting moment is
1.97 HS20 moment. Similar calculations can be performed for other
periods of time. The difference between the mean maximum 50-year
moment and the mean maximum 75-year moment is about 1%.
The mean moments and shears calculated for various time
periods from one day to 75 years are presented in Tables 3.2.4.3-2
and 3.2.4.3-3 for simple moments, shears and negative moments,
respectively. For comparison, the means are also given for an
average truck. All the moments and shears are divided by the
corresponding HS20 moments and shears. The results are also
plotted in Figures 3.2.4.3-4 through 3.2.4.3-6.
The coefficient of variation for the maximum truck moments
and shears can be calculated by transformation of the cumulative
distribution functions (CDF) in Figures 3.2.4.3-1, 3.2.4.3-2 and 3.2.4.3-
3. Each function can be raised to a certain power, so that the
calculated earlier mean maximum moment (or shear) becomes the
mean value after the coefficient of variation. The results are plotted
in Figure 3.2.4.3-7 and Figure 3.2.4.3-8 for moments and shears,
respectively.
printed on June 24, 2003
Lecture - 3-27
Table 3.2.4.3-2 - Mean Maximum Moments for Simple Spans Due to a Single Truck (Divided by
Corresponding HS20 Moment)
Span (m) Average 1 Day 2 Weeks 1 Month 2 Months 6 Months 1 Year 5 Years 50 Years
75 Years
3 0.62 0.97 1.12 1.18 1.23 1.30 1.37 1.46 1.63 1.65
6 0.80 1.30 1.42 1.48 1.53 1.59 1.65 1.76 1.87 1.89
9 0.82 1.33 1.46 1.52 1.57 1.62 1.68 1.78 1.88 1.90
12 0.75 1.31 1.42 1.46 1.50 1.55 1.58 1.64 1.72 1.74
15 0.72 1.32 1.43 1.47 1.52 1.56 1.60 1.65 1.73 1.75
18 0.71 1.37 1.47 1.52 1.56 1.60 1.64 1.69 1.77 1.79
21 0.74 1.42 1.51 1.56 1.60 1.64 1.68 1.74 1.81 1.83
24 0.77 1.47 1.55 1.60 1.64 1.68 1.73 1.79 1.86 1.89
27 0.79 1.51 1.60 1.64 1.68 1.72 1.78 1.84 1.92 1.94
30 0.81 1.55 1.64 1.68 1.72 1.76 1.82 1.89 1.98 2.00
33 0.83 1.60 1.68 1.72 1.76 1.80 1.86 1.94 2.03 2.05
36 0.85 1.63 1.72 1.76 1.80 1.85 1.90 1.97 2.06 2.08
39 0.85 1.66 1.75 1.80 1.83 1.87 1.92 1.99 2.08 2.10
42 0.86 1.67 1.75 1.80 1.83 1.87 1.92 1.99 2.08 2.10
45 0.85 1.64 1.73 1.78 1.81 1.84 1.88 1.96 2.05 2.07
48 0.83 1.60 1.68 1.73 1.76 1.80 1.84 1.91 2.01 2.03
51 0.81 1.55 1.63 1.69 1.72 1.76 1.80 1.87 1.96 1.98
54 0.78 1.50 1.58 1.64 1.67 1.71 1.75 1.82 1.91 1.94
57 0.75 1.45 1.53 1.58 1.62 1.66 1.70 1.77 1.86 1.88
60 0.70 1.38 1.48 1.54 1.57 1.60 1.64 1.71 1.80 1.82
printed on June 24, 2003
Lecture - 3-28
Table 3.2.4.3-3 - Mean Maximum Shears for Simple Spans Due to a Single Truck (Divided by Corresponding
HS20 Shear)
Span Average 1 Day 2 Weeks 1 Month 2 Months 6 Months 1 Year 5 Years 50 Years
75 Years
3 0.78 1.20 1.31 1.38 1.40 1.44 1.48 1.52 1.61 1.62
6 0.72 1.13 1.25 1.29 1.31 1.36 1.38 1.43 1.51 1.52
9 0.68 1.13 1.24 1.29 1.30 1.35 1.38 1.41 1.48 1.49
12 0.66 1.18 1.27 1.32 1.34 1.37 1.40 1.43 1.50 1.51
15 0.69 1.24 1.33 1.37 1.39 1.42 1.45 1.48 1.55 1.56
18 0.73 1.30 1.40 1.44 1.46 1.49 1.51 1.56 1.61 1.62
21 0.74 1.37 1.47 1.50 1.52 1.55 1.58 1.62 1.69 1.70
24 0.77 1.42 1.52 1.57 1.59 1.62 1.65 1.70 1.76 1.77
27 0.80 1.48 1.58 1.62 1.64 1.69 1.72 1.76 1.84 1.85
30 0.81 1.53 1.62 1.66 1.70 1.73 1.76 1.82 1.89 1.90
33 0.82 1.58 1.66 1.70 1.72 1.76 1.80 1.85 1.92 1.93
36 0.83 1.58 1.67 1.71 1.73 1.76 1.80 1.86 1.92 1.93
39 0.83 1.57 1.66 1.70 1.72 1.75 1.77 1.83 1.89 1.90
42 0.81 1.53 1.63 1.66 1.68 1.72 1.74 1.78 1.86 1.87
45 0.79 1.48 1.58 1.62 1.64 1.67 1.70 1.74 1.82 1.83
48 0.76 1.44 1.53 1.57 1.59 1.62 1.65 1.70 1.78 1.79
51 0.74 1.40 1.48 1.52 1.54 1.57 1.60 1.66 1.74 1.75
54 0.71 1.35 1.44 1.47 1.49 1.52 1.56 1.62 1.69 1.70
57 0.69 1.30 1.40 1.43 1.45 1.48 1.51 1.57 1.64 1.65
60 0.68 1.27 1.36 1.39 1.41 1.43 1.47 1.52 1.59 1.60
printed on June 24, 2003
Lecture - 3-29
Table 3.2.4.3-4 - Mean Maximum Negative Moments for Continuous Spans Due to a Single Truck (Divided
by Corresponding HS20 Shear)
Span Average 1 Day 2 Weeks 1 Month 2 Months 6 Months 1 Year 5 Years 50 Years
75 Years
3 0.63 1.12 1.25 1.30 1.33 1.37 1.40 1.46 1.54 1.55
6 0.67 1.30 1.40 1.43 1.44 1.47 1.50 1.53 1.59 1.60
9 0.88 1.50 1.59 1.62 1.64 1.66 1.67 1.72 1.75 1.76
12 0.93 1.63 1.73 1.75 1.77 1.80 1.83 1.86 1.91 1.92
15 0.82 1.50 1.62 1.66 1.67 1.72 1.74 1.78 1.84 1.85
18 0.73 1.34 1.44 1.49 1.51 1.54 1.56 1.61 1.66 1.67
21 0.63 1.24 1.33 1.37 1.39 1.41 1.43 1.47 1.51 1.52
24 0.58 1.16 1.24 1.26 1.29 1.31 1.33 1.35 1.39 1.40
27 0.55 1.11 1.18 1.21 1.22 1.25 1.26 1.29 1.32 1.33
30 0.53 1.07 1.13 1.16 1.17 1.19 1.20 1.22 1.26 1.27
33 0.50 1.03 1.09 1.11 1.12 1.15 1.16 1.18 1.22 1.22
36 0.48 1.00 1.06 1.08 1.09 1.11 1.12 1.15 1.17 1.17
39 0.46 0.96 1.01 1.04 1.05 1.07 1.09 1.10 1.13 1.14
42 0.44 0.94 1.00 1.01 1.02 1.03 1.04 1.07 1.09 1.10
45 0.42 0.90 0.96 0.97 0.99 1.00 1.01 1.03 1.06 1.07
48 0.40 0.86 0.92 0.94 0.95 0.96 0.97 1.00 1.02 1.03
51 0.38 0.84 0.90 0.92 0.93 0.94 0.95 0.97 0.99 1.00
54 0.37 0.81 0.87 0.89 0.89 0.92 0.92 0.94 0.96 0.97
57 0.35 0.80 0.84 0.86 0.87 0.88 0.89 0.91 0.93 0.94
60 0.33 0.78 0.83 0.84 0.85 0.87 0.88 0.89 0.91 0.92
printed on June 24, 2003
Lecture - 3-30
Figure 3.2.4.3-4 - Mean Maximum Moments for Simple Spans Due to
a Single Truck
Figure 3.2.4.3-5 - Mean Maximum Shears for Simple Spans Due to a
Single Truck
printed on June 24, 2003
Lecture - 3-31
Figure 3.2.4.3-6 - Mean Maximum Negative Moments for Two Equal
Continuous Spans Due to a Single Truck
Figure 3.2.4.3-7 - Coefficient of Variation of the Maximum Moment
Due to a Single Truck
printed on June 24, 2003
Lecture - 3-32
Figure 3.2.4.3-8 - Coefficient of Variation of the Maximum Shear Due
to a Single Truck
3.2.4.4 ONE-LANE MOMENTS AND SHEARS
The maximum one-lane moment or shear is caused either by
a single truck or by two or more trucks following each other. For a
multiple truck occurrence, the important parameters are the headway
distance and degree of correlation between truck weights. The
maximum one-lane effect, moment or shear was derived as the largest
of the following two cases:
a. One truck effect, equal to the maximum 75 year
moment or shear with the mean and coefficient of
variation given in Article 3.2.4.3.
b. Two trucks, each with the weight smaller than that of a
single truck above.
Various headway distances were considered from 4.5 to 30 m.
Headway distance is measured from the rear axle of one vehicle to the
front axle of the following vehicle, therefore, 4.5 m means bumper-to-
bumper traffic. Three degrees of correlation between truck weight are
considered: no correlation, partial correlation and full correlation.
It is assumed that, on average, about every 50
th
-100
th
truck is
followed by another truck with the headway distance less than 30 m.
Every 65
th
truck was assumed in this study. In each such
simultaneous occurrence, it is assumed that about every second time
the truck is followed by a partially correlated truck, and about every
10
th
time the truck is followed by a fully correlated truck. The two
trucks are denoted by T
1
and T
2
. The parameters of these two trucks,
including N, i.e., (the considered truck is a maximum of N trucks),
corresponding z = -
-1
(1/N), and T, i.e., (the considered truck is the
maximum for time period T) are given in Table 3.2.4.4-1.
printed on June 24, 2003
Lecture - 3-33
Table 3.2.4.4-1 - Truck Parameters for Two Trucks in One
Lane
One/Two
Trucks
N z T
One 20,000,000 5.33 75 years
Two: No
Correlation
T
1
300,000 4.50 1 year
T
2
1 0.00 average
Partial
Correlation
T
1
150,000 4.36 6 months
T
2
1,000 3.09 1 day
Full
Correlation
T
1
30,000 3.99 1 month
T
2
30,000 3.99 1 month
The maximum values of moments and shears are calculated
by simulations. The parameters considered include truck
configuration, weight, headway distance and frequency of occurrence.
For simple spans, the results of calculations are presented in Figure
3.2.4.4-1 for mean maximum 75 year moments and Figure 3.2.4.4-2
for corresponding shears. For the mean maximum 75 year negative
moments, the results are shown in Figure 3.2.4.4-3. For simple span
moments, one truck governs for spans up to about 42 m for shears up
to about 36 m, and for negative moments in continuous bridges (two
equal spans) up to about 15 m (one span length). The minimum
headway distance is associated with non-moving vehicles or trucks
moving at reduced speeds. This is important in consideration of
dynamic loads.
For simple spans, the calculated mean maximum one-lane
moments are presented in Table 3.2.4.4-2 for time periods from one
day to 75 years. The mean maximum one-lane shears are presented
in Table 3.2.4.4-3. For continuous spans, the mean maximum
negative moments are presented in Table 3.2.4.4-4. The results are
also plotted in Figure 3.2.4.4-4, 3.2.4.4-5 and 3.2.4.4-6.
printed on June 24, 2003
Lecture - 3-34
Table 3.2.4.4-2 - Mean Maximum Moments for Simple Spans Due to Multiple Trucks in One Lane (Divided by Corresponding HS20
Moment)
Span (m) 1 Day 2 Weeks 1 Month 2 Months 6 Months 1 Year 5 Years 50 Years 75 Years
3 0.97 1.12 1.18 1.23 1.30 1.37 1.46 1.65 1.65
6 1.30 1.42 1.48 1.53 1.59 1.65 1.76 1.89 1.89
9 1.33 1.46 1.52 1.57 1.62 1.68 1.78 1.90 1.90
12 1.31 1.42 1.46 1.50 1.55 1.58 1.64 1.74 1.74
15 1.32 1.43 1.47 1.52 1.56 1.60 1.65 1.75 1.75
18 1.37 1.47 1.52 1.56 1.60 1.64 1.69 1.79 1.79
21 1.42 1.51 1.56 1.60 1.64 1.68 1.74 1.83 1.83
24 1.47 1.55 1.60 1.64 1.68 1.73 1.79 1.89 1.89
27 1.51 1.60 1.64 1.68 1.72 1.78 1.84 1.94 1.94
30 1.55 1.64 1.68 1.72 1.76 1.82 1.89 2.00 2.00
33 1.60 1.68 1.72 1.76 1.80 1.86 1.94 2.05 2.05
36 1.63 1.72 1.76 1.80 1.85 1.90 1.97 2.08 2.08
39 1.66 1.75 1.80 1.83 1.87 1.92 1.99 2.10 2.10
42 1.67 1.75 1.80 1.84 1.87 1.92 1.99 2.10 2.10
45 1.66 1.75 1.80 1.83 1.87 1.91 1.99 2.10 2.10
48 1.65 1.74 1.79 1.82 1.85 1.90 1.97 2.08 2.08
51 1.63 1.71 1.77 1.80 1.84 1.88 1.95 2.06 2.06
54 1.60 1.68 1.73 1.77 1.80 1.85 1.91 2.03 2.03
57 1.56 1.65 1.70 1.74 1.78 1.82 1.89 2.00 2.00
60 1.52 1.62 1.66 1.71 1.74 1.78 1.85 1.96 1.96
Table 3.2.4.4-3 - Mean Maximum Shears for Simple Spans Due to Multiple Trucks in One Lane (Divided by Corresponding HS20 Shear)
printed on June 24, 2003
Lecture - 3-35
Span (m) 1 Day 2 Weeks 1 Month 2 Months 6 Months 1 Year 5 Years 50 Years 75 Years
3 1.20 1.31 1.38 1.40 1.44 1.48 1.52 1.61 1.62
6 1.13 1.25 1.29 1.31 1.36 1.38 1.43 1.51 1.52
9 1.13 1.24 1.29 1.30 1.35 1.38 1.41 1.48 1.49
12 1.18 1.27 1.32 1.34 1.37 1.40 1.43 1.50 1.51
15 1.24 1.33 1.37 1.39 1.42 1.45 1.48 1.55 1.56
18 1.30 1.40 1.44 1.46 1.49 1.51 1.56 1.61 1.62
21 1.37 1.47 1.50 1.52 1.55 1.58 1.62 1.69 1.70
24 1.42 1.52 1.57 1.59 1.62 1.65 1.70 1.76 1.77
27 1.48 1.58 1.62 1.64 1.69 1.72 1.76 1.84 1.85
30 1.53 1.62 1.66 1.70 1.73 1.76 1.82 1.89 1.90
33 1.57 1.66 1.70 1.72 1.76 1.80 1.85 1.92 1.93
36 1.59 1.67 1.71 1.73 1.76 1.80 1.86 1.92 1.93
39 1.59 1.67 1.71 1.73 1.76 1.79 1.85 1.91 1.92
42 1.57 1.66 1.70 1.72 1.75 1.77 1.82 1.89 1.90
45 1.53 1.63 1.67 1.69 1.72 1.75 1.79 1.87 1.88
48 1.50 1.59 1.62 1.64 1.67 1.71 1.75 1.84 1.85
51 1.47 1.55 1.60 1.62 1.64 1.67 1.74 1.81 1.82
54 1.44 1.53 1.56 1.58 1.61 1.65 1.71 1.78 1.79
57 1.41 1.50 1.53 1.55 1.58 1.61 1.67 1.74 1.75
60 1.39 1.48 1.51 1.53 1.55 1.59 1.64 1.71 1.72
Table 3.2.4.4-4 - Mean Maximum Negative Moments for Continuous Spans Due to Multiple Trucks in One Lane (Divided by
Corresponding HS20 Negative Moment)
printed on June 24, 2003
Lecture - 3-36
Span (m) 1 Day 2 Weeks 1 Month 2 Months 6 Months 1 Year 5 Years 50 Years 75 Years
3 1.12 1.25 1.30 1.33 1.37 1.40 1.46 1.54 1.55
6 1.30 1.40 1.43 1.44 1.47 1.50 1.54 1.59 1.60
9 1.50 1.59 1.62 1.64 1.66 1.68 1.72 1.75 1.76
12 1.63 1.73 1.75 1.77 1.81 1.83 1.86 1.91 1.92
15 1.58 1.67 1.69 1.71 1.75 1.77 1.80 1.85 1.86
18 1.72 1.83 1.85 1.87 1.92 1.94 1.97 2.02 2.03
21 1.80 1.92 1.94 1.96 2.01 2.03 2.06 2.12 2.13
24 1.80 1.91 1.94 1.96 2.00 2.03 2.06 2.12 2.13
27 1.75 1.86 1.88 1.90 1.95 1.97 2.00 2.06 2.07
30 1.70 1.81 1.83 1.85 1.89 1.91 1.94 2.00 2.01
33 1.66 1.76 1.78 1.80 1.84 1.86 1.89 1.95 1.96
36 1.62 1.72 1.74 1.76 1.80 1.82 1.85 1.90 1.91
39 1.58 1.68 1.70 1.72 1.76 1.78 1.81 1.85 1.86
42 1.55 1.64 1.66 1.68 1.72 1.74 1.77 1.81 1.82
45 1.52 1.61 1.63 1.65 1.69 1.70 1.73 1.78 1.79
48 1.49 1.58 1.60 1.62 1.66 1.67 1.70 1.75 1.76
51 1.46 1.55 1.57 1.59 1.63 1.64 1.67 1.72 1.73
54 1.44 1.53 1.55 1.57 1.60 1.62 1.64 1.69 1.70
57 1.42 1.51 1.52 1.54 1.58 1.59 1.62 1.66 1.67
60 1.42 1.48 1.50 1.52 1.55 1.57 1.60 1.64 1.65
printed on June 24, 2003
Lecture - 3-37
Figure 3.2.4.4-1 - Mean Maximum Moments for Simple Spans Due
to Multiple Trucks in One Lane
Figure 3.2.4.4-2 - Mean Maximum Shears for Simple Spans Due to
Multiple Trucks in One Lane
printed on June 24, 2003
Lecture - 3-38
Figure 3.2.4.4-3 - Mean Maximum Moments for Two Equal Continuous
Spans Due to Multiple Trucks in One Lane
Figure 3.2.4.4-4 - Mean Maximum Moments for Simple Spans Due to
Multiple Trucks in One Lane
printed on June 24, 2003
Lecture - 3-39
Figure 3.2.4.4-5 - Mean Maximum Shears for Simple Spans Due to
Multiple Trucks in One Lane
Figure 3.2.4.4-6 - Mean Maximum Moments for Two Equal Continuous
Spans Due to Multiple Trucks in One Lane
3.2.4.5 GIRDER DISTRIBUTION FACTORS
The analysis of two-lane loading involves the distribution of
truck load to girders. The structural analysis was performed using the
finite element method. The model is based on a linear behavior of
girders and slab. The calculations were performed for spans ranging
from 9 to 60 m, and five cases of girder spacing were considered: 1.2,
1.8, 2.4, 3.0 and 3.6 m. For each case of span and girder spacing,
girder distribution factors were calculated for various truck positions by
moving the truck transversely by 0.3 m at a time. The resulting truck
"influence lines" are used for calculation of the joint effect of two trucks
in adjacent lanes by superposition.
printed on June 24, 2003
Lecture - 3-40
The resulting girder distribution factors (GDF) are compared
with the AASHTO (1989) values and those recommended by Zokaie,
Osterkamp and Imbsen (NCHRP Project 12-26/1).
For moment in an interior girder, AASHTO (1989) specifies a
GDF as follows:
(3.2.4.5-1)
GDF
s
300D
where s is the girder spacing and D is a constant, equal to 5.5 for steel
girders and prestressed concrete girders, and D = 6.0 for reinforced
concrete T-beams. The design moment in a girder is equal to the
product of s/300D and 0.5 of the HS20 moment.
Zokaie, Osterkamp and Imbsen (NCHRP 12-26) proposed
GDF as a function of girder spacing, s (mm), and span length, L (mm).
For interior girders (steel, prestressed concrete and reinforced
concrete T-beams) the formula is
(3.2.4.5-2) GDF 0.15
s
910
0.6
s
L
0.2
For shear, AASHTO (1989) specifies a GDF given by Equation
3.2.4.5-1, except for the axle directly over the support. It is assumed
that over the support, the slab is simply supported by the girders.
Zokaie, Osterkamp and Imbsen (NCHRP Project 12-26/1)
developed the following formula for GDF for shear:
(3.2.4.5-3) GDF 0.4
s
1830
s
7620
2
The results of calculations performed as a part of this study,
along with the GDF's obtained using Equations 3.2.4.5-1, 3.2.4.5-2
and 3.2.4.5-3, are listed in Table 3.2.4.5-1. AASHTO (1989) values
are calculated for steel and prestressed concrete girders using D = 5.5
(denoted by S & P/C in Table 3.2.4.5-1, and for reinforced concrete T-
beams using D = 6 (denoted by R/C in Table 3.2.4.5-1). In Table
3.2.4.5-1, the GDF's calculated in this study are denoted by Nowak,
and those obtained using Equations 3.2.4.5-2 and 3.2.4.5-3 are
denoted by Zokaie, et al. Note that some modifications were made to
the results of NCHRP 12-26 before inclusion in the LRFD Specification
so that the results in Table 3.2.4.5-1 may not be precisely duplicated
using the equations given in the LRFD Specification.
printed on June 24, 2003
Lecture - 3-41
Table 3.2.4.5-1 - Girder Distribution Factors for Interior Girders
Span Girder Moments Shears
Span Spacing AASHTO (1989) Nowak Zokaie AASHTO (1989) Zokaie
(m) (m) S & P/C R/C et al S & P/C RC et al
9 1.2 0.73 0.67 0.88 0.94 0.90 0.88 1.04
9 1.8 1.09 1.00 1.20 1.25 1.25 1.21 1.34
9 2.4 1.45 1.33 1.50 1.53 1.65 1.60 1.63
9 3.0 1.82 1.67 1.79 1.80 1.94 1.88 1.91
9 3.6 2.18 2.00 2.06 2.06 2.28 2.21 2.17
18 1.2 0.73 0.67 0.83 0.84 0.87 0.84 1.04
18 1.8 1.09 1.00 1.10 1.11 1.22 1.17 1.34
18 2.4 1.45 1.33 1.35 1.35 1.61 1.55 1.63
18 3.0 1.82 1.67 1.59 1.59 1.91 1.84 1.91
18 3.6 2.18 2.00 1.82 1.82 2.26 2.18 2.17
27 1.2 0.73 0.67 0.78 0.79 0.86 0.83 1.04
27 1.8 1.09 1.00 1.03 1.03 1.21 1.16 1.34
27 2.4 1.45 1.33 1.26 1.26 1.60 1.54 1.63
27 3.0 1.82 1.67 1.48 1.48 1.91 1.83 1.91
27 3.6 2.18 2.00 1.69 1.69 2.26 2.17 2.17
36 1.2 0.73 0.67 0.73 0.75 0.86 0.83 1.04
36 1.8 1.09 1.00 0.98 0.98 1.21 1.16 1.34
27 2.4 1.45 1.33 1.26 1.26 1.60 1.54 1.63
27 3.0 1.82 1.67 1.48 1.48 1.91 1.83 1.91
27 3.6 2.18 2.00 1.69 1.69 2.26 2.17 2.17
36 1.2 0.73 0.67 0.73 0.75 0.86 0.83 1.04
36 1.8 1.09 1.00 0.98 0.98 1.21 1.16 1.34
36 2.4 1.45 1.33 1.20 1.20 1.60 1.53 1.63
36 3.0 1.82 1.67 1.40 1.40 1.91 1.83 1.91
36 3.6 2.18 2.00 1.60 1.60 2.25 2.16 2.17
60 1.2 0.73 0.67 0.69 0.69 0.75 0.71 1.04
60 1.8 1.09 1.00 0.90 0.90 1.12 1.06 1.34
60 2.4 1.45 1.33 1.10 1.10 1.50 1.41 1.63
60 3.0 1.82 1.67 1.28 1.28 1.87 1.76 1.91
60 3.6 2.18 2.00 1.46 1.46 2.23 2.10 2.17
printed on June 24, 2003
Lecture - 3-42
3.2.4.6 TWO-LANE MOMENTS AND SHEARS
The analysis involves the determination of the load in each
lane and load distribution to girders. The effect of multiple trucks is
calculated by superposition. The maximum moments are calculated
as the largest of the following cases:
1. One lane fully loaded and the other lane unloaded.
2. Both lanes loaded; three degrees of correlation
between the lane loads are considered: no correlation
( = 0), partial correlation ( = 0.5) and full correlation
( = 1).
It has been observed that, on average, about every 10
th
-20
th
(13
th
used) truck is on the bridge simultaneously with another truck
(side-by-side). For each such simultaneous occurrence, it is assumed
that every 10
th
time the trucks are partially correlated and every 30
th
time they are fully correlated (with regard to weight). It is also
assumed that the transverse distance between two side-by-side trucks
is 1.2 m (wheel center-to-center).
The parameters of lane load, including N, i.e, the considered
lane load is the maximum of N occurrences, z =
-1
(1/N), and T, i.e.,
the considered lane load is the maximum in time period T, are given
in Table 3.2.4.6-1.
Table 3.2.4.6-1 - Lane Load Parameters for Two-Lane Traffic
One/Two Lanes Loaded N z T
One 20,000,000 5.33 75 years
Two: = 0 L
1
1,500,000 4.83 5 years
L
2
1 0.00 average
= 0.5 L
1
150,000 4.36 6 months
L
2
1,000 3.09 1 day
= 1 L
1
50,000 4.11 2 months
L
2
50,000 4.11 2 months
The results of simulations indicate that for interior girders, the
case with two fully correlated side-by-side trucks governs, with each
truck equal to the maximum two-month truck. The ratio of a mean
maximum 75-year moment (or shear) and a mean two-month moment
(or shear) is about 0.85 for all the spans.
printed on June 24, 2003
Lecture - 3-43
3.3 LIVE LOADS
3.3.1 Notional Live Load Model
The vehicular live loading consists of a combination of the
design truck or the design tandem, and the design lane load. Each
design lane is occupied by a single design truck or tandem, each
coincident with the lane load, where applicable. The loads shall be
assumed to occupy 3 m transversely.
Generally, the number of design lanes is determined by taking
the integer part of the ratio w/3600, where w is the clear roadway
width between curbs and/or barriers in mm. Fractional design lanes
are not considered. Possible future changes in the physical or
functional clear roadway width of the bridge should be considered. In
cases where the traffic lanes are less than 3600 mm wide, the number
of design lanes is equal to the number of traffic lanes, and the width
of the design lane is taken as the width of the traffic lane. Roadway
widths from 6000 mm to 7200 mm are designed for two design lanes,
each equal to one-half the roadway width.
The axle weights and spacings of the design truck are
indicated in Figure 3.3.1-1. Generally, the spacing between the two
145 kN axles is varied from 4300 mm to 9000 mm to produce extreme
force effects. In exterior lanes or lanes adjacent to median barriers,
the center of the outside wheel is positioned as follows:
For the design of the deck overhang - 300 mm from the face
of the curb or railing
All other components - 600 mm from the face of the curb or
barrier
In all other cases, the center of a wheel need not be placed closer
than 600 mm from the edge of a design lane.
printed on June 24, 2003
Lecture - 3-44
Figure 3.3.1-1 - HS20 Truck
The design tandem consists of a pair of 110 kN axles spaced
1200 mm apart. The transverse spacing of wheels shall be taken as
1800 mm.
The design lane load consists of a load, uniformly distributed
in the longitudinal direction, of 9.3 N/mm. The live load shall be
assumed to occupy 3000 mm transversely.
3.3.2 Multiple Presence of Live Load
Where live load is assigned to components using analysis
methods which require the engineer to explicitly consider the number
of lanes of traffic on the deck, the factors in Table 3.3.2-1 may be used
to reduce calculated extreme force effects in accordance with the
number of lanes loaded for calculating the extreme force effect. This
means that when the multi-lane distribution factors, given in Section
4 of the Specification, are used, instead of the lever rule or statical
assignment, the designer does not know whether two lanes, three
lanes or four lanes, etc. controlled the design because this has been
built into the distribution factor, including the appropriate multiple
factor. Similarly, when a single-loaded lane controls the design for
limit states other than the fatigue limit state, the multiple presence
factor of 1.2 has been included in the single-lane distribution factors.
When the lever rule is used or any other form of statical
assignment of live load, then the designer makes explicit consideration
of the number of loaded lanes. It is common to try one lane with an
appropriate multiple presence factor, two lanes, three lanes and an
appropriate multiple presence factor and continue until a full number
of design lanes is loaded, and then to select a combination of loaded
printed on June 24, 2003
Lecture - 3-45
lanes and multiple presence factor which produce the maximum load
effect on the component under design. This traditional process, where
applicable, would apply using the LRFD Specification, except that the
multiple presence factors themselves are somewhat different, as
indicated in Table 3.3.2-1.
Multiple presence factors are not to be applied to the fatigue
limit state for which one design truck is used, regardless of the number
of design lanes, which is almost always one lane loaded. Thus, the
factor 1.20 must be removed from the single lane distribution factors
when they are used to investigate fatigue.
Table 3.3.2-1 - Multiple Presence Factors "m"
Number of
Design Lanes
Multiple Presence
Factors "m"
1 1.20
2 1.00
3 0.85
>3 0.65
Additional guidance is provided for bridges which support
roadways and sidewalks.
3.3.3 Application of Design Vehicular Live Loads
Generally, the extreme force effect is to be taken as the larger
value determined from the following loading combinations:
effect of the design tandem combined with the effect of the
design lane load,
effect of one design truck combined with the effect of the
design lane load, and
for negative moment between points of dead load
contraflexure and reaction at interior piers only, 90% of the
effect of two design trucks spaced a minimum of 15 m
between the rear axle of one truck and the front axle of the
following truck, and 4.3 m between the two 145 kN axles,
combined with 90% of the effect of the design lane load.
Except where noted, the design truck or design tandem shall
be taken as coincident with the lane design load. The lane load is not
interrupted to provide space for the axle sequences of the design
tandem or the design truck; interruption is needed only for patch
loading patterns to produce extreme force effects. Axles which do not
contribute to the extreme force effect under consideration shall be
neglected. The lane load is positioned longitudinally for extreme
effect.
printed on June 24, 2003
Lecture - 3-46
In order to calculate force effect for a component, design lanes
and the position of the 3 m loaded width in each lane are be
positioned on the bridge deck to produce extreme force effects.
Design lanes, or parts thereof, which contribute to the extreme force
effect under consideration, are loaded.
3.3.4 Fatigue Requirements
In the case of the fatigue limit state, a single design truck, or
one or more axles thereof, is specified as the live load to be
considered. Note that neither the tandem axle pair or the design lane
load is applied in this limit state. Furthermore, the distance between
the two large axles on the truck is a constant 9 m, rather than the
variable spacing used for strength design. The reason for this
simplification is that the majority, and, therefore, the statistically
significant number of vehicles on the road are relatively long 3S-2
configurations, and it would be unduly severe to assume that all of the
fatigue stress ranges result from the smaller number of relatively short
trucks. In addition to the size of the vehicle, the number of single-lane
occurrences is necessary in order to predict the number of cycles of
fatigue loading. This is referred to in the Specification as a single-lane
average daily truck traffic, and in the absence of site or road specific
information, may be related to the typically tabulated average daily
truck traffic volume through a percentage factor, specified in the
Specification, to range from 100% for a single-lane available for truck
traffic to as low as 80% when three or more lanes are available.
Additionally, where only average daily traffic counts are available, the
commentary to Article S3.6.1.4.2 provides factors which may be used
to relate ADT to ADTT. These factors have been developed based on
observations of traffic counts throughout the Country and have to be
regarded as approximate, but will serve the purpose when no other
data is available.
3.3.5 Tire Pressure
Where the weight of a wheel, assumed to consist of two
closely spaced tires, is considered as a contact load, the length of the
tire patch may be assumed to be 510 mm in width and whose length
is given by Equation S3.6.1.2.5-1, which is repeated below.
= 2.28 (1 + IM/100) P (3.3.5-1)
Note that the assumed tire contact area is assumed to increase with
the load factor and the impact. This is because the tire pressure has
been assumed to be a constant 0.862 MPa, a value which is
consistent with on-the-road tires.
Figure 3.3.5-1, Musser, 1992, shows data on the pressure
taken from an inventory of actual tire load, pressure and footprint
characteristics based on detailed engineering data provided by
industry sources. This data validates the use of 0.862 MPa.
printed on June 24, 2003
Lecture - 3-47
Higher pressures can be generated by off-the-road tires, but
that factor is not included in the LRFD Specification.
Additional guidance on how to distribute the tire pressure on
uniform and interrupted services is presented in the Specification.
Figure 3.3.5-1 - Tire Pressures Based on Industry Data
3.3.6 Live Load Deflection Criteria
When calculating live load deflection for the optional deflection
provisions in Article S2.5.2.6.2, it is specified that deflection should be
the larger of that resulting from the truck alone or that resulting from
one-quarter of the truck, plus the uniform load. These two criteria
were established to be consistent with the type of loading for which
deflection was calculated under previous additions of the Standard
Specification. It is generally thought that the entire issue of deflections
is obscure enough that it was not necessary to require more load than
that for which previous calculations would have been developed.
3.3.7 Dynamic Load Allowance
Past editions of the AASHTO Specification provided for the
increase in apparent magnitude of force effects of live load due to the
dynamic response of the bridge itself. Traditionally, the extent of
dynamic load amplification was considered to be solely a function of
the span length. Early editions of the Ontario Highway Bridge Design
Specification tried to relate the dynamic load effect to a frequency of
the superstructure. Research has shown that the interaction of the
weight and configuration of the vehicle, the springs and damping
characteristics provided by springs and tires, as well as the surface
condition of the roadway, are important variables. Generally speaking,
printed on June 24, 2003
Lecture - 3-48
the dynamic amplification of trucks follows the following general
trends:
As the weight of the vehicle goes up, the apparent
amplification goes down.
Multiple vehicles produce a lower dynamic amplification than
a single vehicle.
More axles result in lower dynamic load amplification.
A study of dynamic effects was reported in the report of the
Calibration Task Group, Nowak, 1993, from which the following figures
are taken. The method of evaluating dynamic effects was to compare
static deflection to deflection under simulated moving live load. The
dynamic load factor (DLF), was defined as the maximum dynamic
deflection, D
dyn
, divided by the maximum static deflection, D
sta
, as
shown in Figure 3.3.7-1.
Figure 3.3.7-1 - Time History for a Typical Bridge Mid-Span Deflection
The cross-sections studied are shown in Figure 3.3.7-2.
printed on June 24, 2003
Lecture - 3-49
Figure 3.3.7-2 - Cross-Sections for Bridges Considered in Simulations
The effect of vehicle weight for a five-axle truck and a three-axle truck
on the steel girder cross-section for various span lengths is shown in
Figures 3.3.7-3 and 3.3.7-4, respectively.
Figure 3.3.7-3 - Dynamic Load Effect vs. Gross Vehicle Weight for a
Five-Axle Truck on Steel Girder Bridge
printed on June 24, 2003
Lecture - 3-50
Figure 3.3.7-4 - Dynamic Load Effect vs. Gross Vehicle Weight for a
Three-Axle Truck on Steel Girder Bridge
Comparable results for a five-axle truck on prestressed
concrete beam cross-sections on various span bridges are shown in
Figure 3.3.7-5.
Figure 3.3.7-5 - Dynamic Load vs. Gross Vehicle Weight for a Five-
Axle Truck on Prestressed Concrete Girder Bridges
In developing the data shown in Figures 3.3.7-3 through 3.3.7-
5, it was observed that the dynamic portion of the deflection was not
very much effected by gross vehicle weight, whereas the static portion
of the total deflection was. Therefore, the apparent reduction in
dynamic effect, shown as the weight of the vehicle increases, is
primarily due to the increased static deflection.
It is commonly thought that vehicle velocity is a primary factor
in dynamic effects. Figure 3.3.7-6 shows the effect of increasing
vehicle speed for three different span lengths of steel girder bridges.
Some decrease in dynamic effect with increase of speed is apparent
for heavier vehicles, but not necessarily for lighter vehicles. The effect
of road roughness on dynamic effects for various weights of vehicle
are shown in Figure 3.3.7-7, the increase in dynamic effect with
increasing roughness is evident for a wide range of vehicle weights.
printed on June 24, 2003
Lecture - 3-51
Figure 3.3.7-6 - Dynamic Load Effect vs. Truck Speed for Five-Axle
Trucks on Steel Girder Bridges
Figure 3.3.7-7 - Dynamic Load vs. Roughness Coefficient
The influence of a second vehicle on the structure is shown on
a basis of time history in Figure 3.3.7-8 and as a dynamic load factor
in Figure 3.3.7-9. The latter figure shows average values from a
simulation of many combinations of roughness and suspension
characteristics. The generally beneficial effect of multiple presence of
vehicles is clearly indicated.
printed on June 24, 2003
Lecture - 3-52
Figure 3.3.7-8 - Time History for Two Trucks
Figure 3.3.7-9 - Average Dynamic Load vs. Span
After consideration of the study, cited above, and other studies
of dynamic load, the requirements outlined in Article S3.6.2.1 and
summarized below, were developed.
printed on June 24, 2003
Lecture - 3-53
Component IM
Deck Joints - All Limit States 75%
All Other Components
! Fatigue and Fracture
Limit State
! All Other Limit States
15%
33%
The provisions required that the general dynamic load
allowance would be taken as 1/3 of the weight of the design truck, with
no dynamic load allowance applied to the design uniform load. The
factor of 1/3, or 33-1/3%, was selected from two considerations. The
starting point was an amplification of 25%, which is reasonably
indicative of the results cited above for a 325 kN vehicle. However,
the purpose of superimposing the uniform load and the truck in a
combined load model is to recognize that a single vehicle in a lane
may weigh more than 325 kN. As the length of the span increases,
then the uniform load serves two purposes, not only simulation of
heavier single vehicles in the lane, but the presence of other traffic.
Since increased traffic tends to dissipate the dynamic effect, the
decision was made to increase the dynamic load allowance from 25%
to 33-1/3% when the load model was simulating the effect of single,
heavy trucks, but to consider no dynamic influence from the remainder
of the uniform load as the span increases. It could be argued that the
figures presented above would indicate that no increase in dynamic
effect would be necessary as the weight of the vehicle increased, but
the decision was made to treat the subject somewhat conservatively
because of the departure from previous specifications.
The requirement for 75% impact on deck joints is based on
oral reports of tests in Europe in which parts of expansion joints were
strain-gaged and were found to be subject to about 75% dynamic load
increase, as well as the observed tendency for there to be an even
settlement immediately preceding bridges, as well as dislodged,
broken concrete in the immediate vicinity of expansion dams, all of
which served to increase the impact.
Some relief from the dynamic load allowance provision may be
obtained if rigorous dynamic analyses are provided to support the
reduction. This is not expected to be the case for typical designs, but
may be of value when larger bridges, and/or repetitive viaduct
structures are being designed.
Drafts of the LRFD Specification, developed prior to adoption,
considered the possibility of providing a separate load allowance for
decks and deck systems. While there is some technical merit in doing
this, that provision produced considerable confusion during the trial
printed on June 24, 2003
Lecture - 3-54
designs, and was somewhat ambiguous in its application of deck
systems which support local and global loads. For simplicity, the
distinction between decks and other parts of the superstructure was
removed from the Specification.
The reduced dynamic allowance for the fatigue limit state is
based on the premise that average, rather than extreme values, are
more appropriate when considering cumulative damage from fatigue.
Article S3.6.2 permits the traditional elimination of dynamic
load allowance for retaining walls and components of structures which
are below ground, given below, and provides for a variable increase
in dynamic load allowance for culverts and other buried structures,
depending on the depth of earth fill.
0% IM 40 (1.0 4.1x10
4
D
E
)
where:
D
E
= the minimum depth of earth cover above the structure
(mm)
Wood structures typically experience a great deal of damping
which tends to reduce dynamic load amplification. Accordingly, they
have been allowed a 50% reduction in the standard dynamic load
allowance values.
3.3.8 Miscellaneous Live Loads
The Specification recognizes that pedestrian loads are also
significant and provisions for both pedestrian and bicycle traffic is
provided.
Similarly, it is anticipated that more bridges will have rail transit
loads in the future and the need to consider that is highlighted.
Approximations are provided for deck overhang loads to
simplify calculations by replacing the design load with a line load.
Q1-1
QUIZ
Given: A simple span girder of 30 m span from center-to-center of
bearings
Find: a) centerline moment from live load, plus dynamic load
allowance
b) reaction from live load, plus dynamic load allowance
Solution
UDL:
w x L
2
8
9.3 x 30 000 x 30 000
8
1.046 x 10
9
N mm
Truck:
15 000 4300
15 000
(7500) (35 000 145 000)
7500 (145 000) 2.051 x 10
9
N mm
Tandem:
15 000 1200
15 000
1 (7500) (110 000) 1.584 x 10
9
N mm
Truck > Tandem
M
a
(2.051 x 1.33 1.046) x 10
9
3.774 x 10
9
N mm
Q1-2
UDL
9.3 x 30 000
2
1.395 x 10
5
N
Truck 1
30 000 4300
30 000
(145 000)
30 000 8600
30 000
(35 000) 2.942 x 10
5
N
Tandem 1
30 000 1200
30 000
(110 000) 2.156 x 10
5
N
Truck > Tandem
Reaction (2.942 x 1.33 1.395) 10
5
5.308 x 10
5
N
printed on June 24, 2003
Work Period #1 - 1
WORK PERIOD #1: Live Loads on Multi-Span Bridges
FHWA COURSE ON LRFD DESIGN OF HIGHWAY BRIDGES
LRFD SHEARS AND MOMENTS PER LANE - MUST BE DISTRIBUTED TO A GIRDER
DYNAMIC LOAD ALLOWANCE HAS NOT BEEN ADDED EXCEPT WHERE INDICATED
LRFD SPECIFICATION (SI UNITS)
SINGLE TRUCK = 1053.9 kN.m
TWO TRUCKS = 1412.8 kN.m
TANDEM = 724.2 kN.m
UNIFORM = 859.8 kN.m
FATIGUE TRUCK = 1005.8 kN.m
NEGATIVE LL MOMENT (INCLUDING DYNAMIC LOAD ALLOWANCE):
SINGLE TRUCK + UNIFORM = 1053.9*1.33+859.8 = 2261.5 kN.m
TWO TRUCKS + UNIFORM = 0.9*(1412.8*1.33+859.8) = 2464.9 kN.m
TANDEM + UNIFORM = 724.2*1.33+859.8 = 1822.9 kN.m
FATIGUE TRUCK = 1005.8*1.15 = 1156.7 kN.m
Note: In case of two trucks, the distance between the rear axle of the
leading truck and the front axle of the trailing truck required
to produce maximum effect is 15.0 m.
printed on June 24, 2003
Work Period #1 - 2
POSITIVE LL MOMENT AT 0.8L
==========================
SINGLE TRUCK = 1135.1 kN.m
TANDEM = 924.2 kN.m
UNIFORM = 515.9 kN.m
FATIGUE TRUCK = 966.1 kN.m
POSITIVE LL MOMENT (INCLUDING DYNAMIC LOAD ALLOWANCE):
SINGLE TRUCK +UNIFORM = 1135.1*1.33+515.9 = 2025.6 kN.m
TANDEM + UNIFORM = 924.2*1.33+515.9 = 1745.1 kN.m
FATIGUE TRUCK = 966.1*1.15 = 1111.1 kN.m
printed on June 24, 2003
Work Period #1 - 3
POSITIVE LL MOMENT AT 0.4L
==========================
SINGLE TRUCK = 2529.9 kN.m
TANDEM = 1894.4 kN.m
UNIFORM = 1633.6 kN.m
FATIGUE TRUCK = 2253.1 kN.m
POSITIVE LL MOMENT (INCLUDING DYNAMIC LOAD ALLOWANCE):
SINGLE TRUCK + UNIFORM = 2529.9*1.33+1633.6 = 4998.3 kN.m
TANDEM + UNIFORM = 1894.4*1.33+1633.6 = 4153.1 kN.m
FATIGUE TRUCK = 2253.1*1.15 = 2591.0 kN.m
printed on June 24, 2003
Work Period #1 - 4
SINGLE TRUCK = 527.0 kN.m
TANDEM = 361.8 kN.m
UNIFORM = 412.7 kN.m
FATIGUE TRUCK = 503.2 kN.m
NEGATIVE LL MOMENT (INCLUDING DYNAMIC LOAD ALLOWANCE):
SINGLE TRUCK + UNIFORM = 527.0*1.33+412.7 = 1113.5 kN.m
TANDEM + UNIFORM = 361.8*1.33+412.7 = 893.9 kN.m
FATIGUE TRUCK = 503.2*1.15 = 578.7 kN.m
printed on June 24, 2003
Work Period #1 - 5
SINGLE TRUCK = 1317.8 kN.m
TWO TRUCKS = 2627.6 kN.m
TANDEM = 905.3 kN.m
UNIFORM = 2149.5 kN.m
FATIGUE TRUCK = 1258.1 kN.m
NEGATIVE LL MOMENT (INCLUDING DYNAMIC LOAD ALLOWANCE):
SINGLE TRUCK + UNIFORM = 1317.8*1.33+2149.5 = 3902.2
kN.m
TWO TRUCKS + UNIFORM = 0.9*(2627.6*1.33+2149.5) = 5079.8 kN.m
TANDEM + UNIFORM = 905.3*1.33+2149.5 = 3353.6 kN.m
FATIGUE TRUCK = 1258.1*1.15 = 1446.8
kN.m
Note: In case of two trucks, the distance between the rear axle
of the leading truck and the front axle of the trailing
truck required to produce maximum effect is 30.1 m.
printed on June 24, 2003
Work Period #1 - 6
SINGLE TRUCK = 322.4 kN
TWO TRUCKS = 579.3 kN
TANDEM = 219.6 kN
UNIFORM = 499.9 kN
LL REACTION (INCLUDING DYNAMIC LOAD ALLOWANCE):
SINGLE TRUCK + UNIFORM = 322.4*1.33+499.9 = 928.7 kN
TWO TRUCKS + UNIFORM = 0.9*(579.3*1.33+499.9) = 1143.3 kN
TANDEM + UNIFORM = 219.6*1.33+499.9 = 791.9 kN
Note: In case of two trucks, the distance between the rear axle of the
leading truck and the front axle of the trailing truck required
to produce maximum effect is 15.0 m.
printed on June 24, 2003
Work Period #1 - 7
POSITIVE REACTION AT LEFT SUPPORT
=================================
SINGLE TRUCK = 298.2 kN
TANDEM = 216.2 kN
UNIFORM = 175.0 kN
FATIGUE TRUCK = 274.2 kN
REACTION (INCLUDING DYNAMIC LOAD ALLOWANCE):
SINGLE TRUCK + UNIFORM = 298.2*1.33+175.0 = 571.7
kN
TANDEM + UNIFORM = 216.2*1.33+175.0 = 462.5 kN
FATIGUE TRUCK = 274.2*1.15 = 315.3
kN
printed on June 24, 2003
Work Period #1 - 8
NEGATIVE REACTION AT LEFT SUPPORT
=================================
SINGLE TRUCK = 30.6 kN
TANDEM = 21.1 kN
UNIFORM = 25.0 kN
FATIGUE TRUCK = 29.3 KN
NEGATIVE REACTION (INCLUDING DYNAMIC LOAD ALLOWANCE):
SINGLE TRUCK + UNIFORM = 30.6*1.33+25.0 = 65.7 kN
TANDEM + UNIFORM = 21.1*1.33+25.0 = 53.0 kN
FATIGUE TRUCK = 29.3*1.15 = 33.6
kN
printed on June 24, 2003
Work Period #1 - 9
SHEAR TO THE LEFT OF INTERMEDIATE SUPPORT
=========================================
SINGLE TRUCK = 312.2 kN
TANDEM = 218.2 kN
UNIFORM = 249.9 kN
FATIGUE TRUCK = 297.6 kN
SHEAR (INCLUDING DYNAMIC LOAD ALLOWANCE):
SINGLE TRUCK + UNIFORM = 312.2*1.33+249.9 = 665.1
kN
TANDEM + UNIFORM = 218.2*1.33+249.9 = 540.1 kN
FATIGUE TRUCK = 249.5*1.15
= 342.3 kN
printed on June 24, 2003
Work Period #1 - 10
FHWA COURSE ON LRFD DESIGN OF HIGHWAY BRIDGES
LRFD SHEARS AND MOMENTS PER LANE - MUST BE DISTRIBUTED TO A GIRDER
DYNAMIC LOAD ALLOWANCE HAS BEEN ADDED
LRFD SPECIFICATION (SI UNITS)
EI DATA FOR SPAN 1 , LENGTH = 43.00
DIST EI
0.00 1.00
43.00 1.00
EI DATA FOR SPAN 2 , LENGTH = 43.00
DIST EI
0.00 1.00
43.00 1.00
WEIGHT OF ONE AXLE OF DUAL AXLE LOADING 110.00 kN
PROPOSED AASHTO LIVE LOAD ENVELOP - BEFORE LATERAL
DISTRIBUTION
M O M E N T - kN-m S H E A R -
kN
SPAN X-IL +FATG +STREN -FATG -STREN +FATG +STREN -FATG
-STREN
1 0.00 0 0 0 0 315 573 -34
-66
1 4.30 1160 2147 -145 -282 270 482 -34
-68
1 8.60 1941 3670 -290 -564 226 397 -46
-107
1 12.90 2430 4600 -435 -846 183 318 -80
-172
1 17.20 2591 5010 -579 -1128 144 247 -125
-240
1 21.50 2524 4924 -724 -1410 107 183 -168
-311
1 25.80 2292 4390 -869 -1692 73 127 -209
-383
1 30.10 1802 3404 -1014 -1974 44 80 -248
-456
1 34.40 1111 2038 -1159 -2458 24 41 -283
-528
1 38.70 396 695 -1304 -3127 10 16 -315
-598
1 43.00 0 0 -1448 -5064 0 0 -342
-665
2 0.00 0 0 -1448 -5064 342 666 0
0
2 4.30 396 695 -1304 -3127 315 598 -10
-16
2 8.60 1111 2038 -1159 -2458 283 528 -24
-41
2 12.90 1802 3404 -1014 -1974 248 456 -44
printed on June 24, 2003
Work Period #1 - 11
-79
2 17.20 2292 4390 -869 -1692 209 383 -73
-127
2 21.50 2524 4924 -724 -1410 168 311 -107
-183
2 25.80 2591 5010 -579 -1128 125 240 -143
-247
2 30.10 2430 4600 -435 -846 80 172 -183
-318
2 34.40 1941 3670 -290 -564 46 107 -226
-397
2 38.70 1160 2147 -145 -282 34 68 -270
-482
2 43.00 0 0 0 0 34 66 -315
-573
TABLE OF LIVE LOAD REACTIONS - kN
SUPT +T +L -T -L
1 315 573 -34 -66
2 366 1143 0 0
3 315 573 -34 -66
printed on June 24, 2003
Work Period #1 - 12
FHWA COURSE ON LRFD DESIGN OF HIGHWAY BRIDGES
LRFD SHEARS AND MOMENTS PER LANE - MUST BE DISTRIBUTED TO A GIRDER
DYNAMIC LOAD ALLOWANCE HAS NOT BEEN ADDED
1992 AASHTO (SI UNITS)
EI DATA FOR SPAN 1 , LENGTH = 43.00
DIST EI
0.00 1.00
43.00 1.00
EI DATA FOR SPAN 2 , LENGTH = 43.00
DIST EI
0.00 1.00
43.00 1.00
WEIGHT OF ONE AXLE OF DUAL AXLE LOADING 110 kN
M O M E N T - kN-m S H E A R -
kN
SPAN X-IL +T +L -T -L +T +L -T
-L
1 0.00 0 0 0 0 294 292 -30
-36
1 4.30 1094 972 -130 -140 254 240 -30
-42
1 8.60 1853 1687 -260 -279 215 193 -53
-64
1 12.90 2302 2146 -389 -420 178 151 -93
-90
1 17.20 2495 2356 -519 -560 142 114 -130
-120
1 21.50 2446 2324 -649 -700 109 83 -167
-155
1 25.80 2192 2059 -780 -839 78 57 -201
-192
1 30.10 1733 1574 -908 -980 50 36 -233
-233
1 34.40 1118 880 -1039 -1120 27 20 -262
-276
1 38.70 426 287 -1169 -1677 11 8 -286
-321
1 43.00 0 0 -1299 -2798 0 0 -308
-366
2 0.00 0 0 -1299 -2798 308 366
0 0
2 4.30 426 287 -1169 -1677 286 321 -11
-8
2 8.60 1118 880 -1039 -1120 262 276 -27
-20
2 12.90 1733 1574 -908 -980 233 233 -50
-36
printed on June 24, 2003
Work Period #1 - 13
2 17.20 2192 2059 -780 -839 201 192 -78
-57
2 21.50 2446 2324 -649 -700 167 155 -109
-83
2 25.80 2495 2356 -519 -560 130 120 -142
-114
2 30.10 2302 2146 -389 -420 93 90 -178
-151
2 34.40 1853 1687 -260 -279 53 64 -215
-193
2 38.70 1094 972 -130 -140 30 42 -254
-240
2 43.00 0 0 0 0 30 36 -294
-292

TABLE OF LIVE LOAD REACTIONS - kN
SUPT +T +L -T -L
1 294 292 -30 -36
2 318 617 0 0
3 294 292 -30 -36
printed on June 24, 2003
Work Period #1 - 14
FHWA COURSE ON LRFD DESIGN OF HIGHWAY BRIDGES
LRFD SHEARS AND MOMENTS PER LANE - MUST BE DISTRIBUTED TO A GIRDER
DYNAMIC LOAD ALLOWANCE HAS NOT BEEN ADDED
LRFD SPECIFICATION (SI UNITS)
EI DATA FOR SPAN 1 , LENGTH = 43.00
DIST EI
0.00 1.00
43.00 1.00
EI DATA FOR SPAN 2 , LENGTH = 43.00
DIST EI
0.00 1.00
43.00 1.00
WEIGHT OF ONE AXLE OF DUAL AXLE LOADING 110.00 kN
PROPOSED AASHTO LIVE LOAD ENVELOP - BEFORE LATERAL
DISTRIBUTION
M O M E N T - kN-m S H E A R -
kN
SPAN X-IL +FATG +STREN -FATG -STREN +FATG +STREN -FATG
-STREN
1 0 0 0 0 0 274 473 -29
-55
1 4.30 1009 1777 -126 -238 235 396 -29
-58
1 8.60 1688 3044 -252 -476 196 324 -40
-89
1 12.90 2113 3822 -378 -714 160 258 -70
-141
1 17.20 2253 4168 -504 -953 125 199 -108
-196
1 21.50 2195 4098 -630 -1191 93 147 -146
-255
1 25.80 1993 3649 -756 -1429 64 101 -182
-315
1 30.10 1567 2819 -882 -1667 38 63 -215
-377
1 34.40 966 1660 -1008 -2036 21 32 -246
-439
1 38.70 344 549 -1134 -2627 8 12 -274
-501
1 43.00 0 0 -1260 -4277 0 0 -297
-561
2 0 0 0 -1260 -4277 298 562
0 0
2 4.30 344 549 -1134 -2627 274 501 -8
12
2 8.60 966 1660 -1008 -2036 246 440 -21
-32
printed on June 24, 2003
Work Period #1 - 15
2 12.90 1567 2819 -882 -1667 216 377 -38
-63
2 17.20 1993 3649 -756 -1429 182 315 -64
-101
2 21.50 2195 4098 -630 -1191 146 255 -93
-147
2 25.80 2253 4168 -504 -953 108 196 -125
-199
2 30.10 2113 3822 -378 -714 70 141 -160
-258
2 34.40 1688 3044 -252 -476 40 89 -196
-324
2 38.70 1009 1777 -126 -238 29 58 -235
-396
2 43.00 0 0 0 0 29 55 -274
-473
TABLE OF LIVE LOAD REACTIONS - kN
SUPT +SERV +STREN -SERV -STREN
1 274 473 -29 -55
2 318 969 0 0
3 274 473 -29 -55
printed on June 24, 2003
Work Period #1 - 16
FHWA COURSE ON LRFD DESIGN OF HIGHWAY BRIDGES
LRFD SHEARS AND MOMENTS PER LANE - MUST BE DISTRIBUTED TO A GIRDER
DYNAMIC LOAD ALLOWANCE HAS NOT BEEN ADDED
RATIO OF LRFD TO 1992 AASHTO
MOMENT SHEAR
SPAN X-IL +STREN -STREN +STREN -STREN
1 0.00 1.609 1.528
1 4.30 1.624 1.700 1.559 1.381
1 8.60 1.643 1.706 1.507 1.391
1 12.90 1.660 1.699 1.449 1.566
1 17.20 1.671 1.702 1.401 1.633
1 21.50 1.675 1.702 1.349 1.645
1 25.80 1.665 1.703 1.295 1.640
1 30.10 1.627 1.701 1.260 1.617
1 34.40 1.485 1.818 1.185 1.593
1 38.70 1.289 1.566 1.091 1.563
1 43.00 1.528 1.534
2 0.00 1.528 1.534
2 4.30 1.289 1.566 1.563 1.091
2 8.60 1.485 1.818 1.594 1.185
2 12.90 1.627 1.701 1.618 1.260
2 17.20 1.665 1.703 1.641 1.295
2 21.50 1.675 1.702 1.646 1.349
2 25.80 1.671 1.702 1.635 1.401
2 30.10 1.660 1.699 1.566 1.449
2 34.40 1.643 1.706 1.391 1.507
2 38.70 1.624 1.700 1.381 1.559
2 43.00 1.528 1.609

SUPT +STREN -STREN
1 1.619 1.528
2 1.571
3 1.619 1.528

Вам также может понравиться