Sei sulla pagina 1di 300

AREMA

Bridge Inspection
Handbook ©

Developed by
Committee 10
Structures, Maintenance & Construction

AMERICAN RAILWAY ENGINEERING AND MAINTENANCE-OF-WAY ASSOCIATION


The contents of this Handbook are published as a guide to establishing policies and
practices relative to bridge inspection. Actual policies and practices are the prerogative
of individual railroads or their authorized representative(s) based on the nature and
characteristics of their physical plant and operations, as well as the specific
characteristics of the geographical regions through which they operate.

While every effort has been made to ensure its accuracy, nothing in this Handbook
relieves an inspector or supervisor of the responsibility to conform to sound engineering
practice and to conform to any or all local, state, or federal rules or regulations. Anyone
using this material does so at his own risk and assumes any and all liability resulting from
such use. No liability can be accepted by the authors or publishers for loss, damage, or
injury caused by any errors in, or omissions from, the information provided.

Published by
The American Railway Engineering
and Maintenance of Way Association
10003 Derekwood Lane, Suite 210
Lanham, MD 20706-4362 USA
(301) 459-3200
fax (301) 459-8077
www.arema.org

Copyright © 2008 by The American Railway Engineering


and Maintenance of Way Association
All rights reserved.

No part of this publication may be produced, stored in any information


or data retrieval system, or transmitted, in any form, or by any
means – electronic, mechanical, photocopying, recording, or otherwise –
without the prior written permission of the publisher.

ii
CONTRIBUTORS TO THE BRIDGE INSPECTION HANDBOOK

This Handbook is a product of AREMA Committee 10, Structures, Maintenance &


Construction. The principal authors (and AREMA Members) include:

Willie Benton, III


Gordon A. Davids, PE
David K. Franz, PE
Bob Grace
Glenn R. Green
William J. Hager
Bernard L. Hopfinger
John J. Horney, PE
Peter Schierloh
Donald F. Sorgenfrei, PE

In addition, Committee 10 would like to thank the following individuals or groups within
AREMA for their significant, additional contribution to this handbook:

Stephen J. Hill, PE
Richard W. “Dutch” Miller
Peter Schierloh
Committee 9 – Seismic Design for Railway Structures

Other members of Committee 10 participated in various review and editing roles and
their collective effort was very important in the production of this Handbook.
Additionally, Committee 10 thanks the members of AREMA Committees 7, 8, 9, and 15,
as well as the Structures Functional Group Vice President and Directors for their
assistance in reviewing this material.

iii
AREMA BRIDGE INSPECTION HANDBOOK©
TABLE OF CONTENTS

1 Bridge Inspection Practices


General
Safety Practices
Personal Protection
Accident Prevention
Safety Precautions
Climbing Safety
Confined Spaces
Standard Tools
Methods of Access
Access Equipment
2 Confined Spaces
General
Safety
Culverts
Tunnels
Movable Bridges and Turntables
Other Structures
3 Site Conditions
General
Utilities
Adjacent Property Development
Vegetation Growth
Erosion
Adjacent or Parallel Bridge/Culvert Structures
Track/Bridge Geometry
ROW Fences
Vandalism
Site Conditions Checklist
4 Bridge Loads and Forces
General
Types of Loads and Forces
Fatigue
Load Paths
Load Redistribution
Deflection
Bridge Ratings
Other Factors
5 Bridge Nomenclature
General
Substructure Components and Terminology
Superstructure Types, Components, and Terminology
Movable Bridges
Turntables
Transfer Tables
Identifying Structures and Components
Condition Ratings
6 Bridge Decks
General
Deck Defects
Open Deck Bridges
Ballast Deck Bridges
Direct Fixation Deck Bridges
Inner Guard Rails
Deck Inspection Checklist
7 Timber Bridges
General
Material Properties
Bridge Notes
Modes of Timber Deterioration
Inspection Methodology
Details of Inspection
Effects of Unit Trains on Open Deck Timber Bridges
Timber Bridge Inspection Checklist
8 Concrete & Masonry Bridges and Foundations
Section 1 – Concrete Properties
General
Material Mechanics
Engineering Characteristics
Behavior of Reinforced Concrete
Concrete Deterioration
Section 2 – Inspection of Concrete & Masonry Structures and Foundations
General
Underwater Inspections
Foundations
Inspections for Foundation Problems
Substructure Types and Problems
Indications of Substructure Problems
Inspection of Abutments and Piers
Retaining Walls
Bridge Seats and Bearing Areas
Inspection of Bridge Seats and Bearing Areas
Superstructure
Inspection of Concrete Superstructures
Arches
Fire Damage
Concrete Bridge Inspection Checklist
9 Steel Bridges
Section 1 – Iron and Steel Bridges
General
History
Steel – Material
Steel – Fabricated Members
Steel – Fatigue
Steel – Fracture Critical Bridge Members
Steel – Deterioration
Section 2 – Steel - Fastenings
General
Rivets
Bolts
Welds
Section 3 – Bearings
General
Bearings - Materials
Expansion Bearings
Fixed Bearings
Bearing Inspection Checklist
Section 4 – Steel Beam Spans
General
Features
Commonly Found Defects
Inspection Checklist
Section 5 – Floor Systems
General
Types of Floor Systems
Floor System Inspection Checklist
Commonly Found Defects
Section 6 – Through - Plate Girder (TPG) Spans
General
Nomenclature
Features
TPG Inspection
TPG Inspection Checklist
Common TPG Span Defects
Section 7 – Deck Plate Girder (DPG) Spans
General
Nomenclature
Features
Inspection
DPG Inspection Checklist
Commonly Found Defects
Section 8 – Trestles and Viaducts
General
Nomenclature
Features
Trestle and Tower Inspection Checklist
Common Bent and Tower Defects
Section 9 – Trusses
General
Nomenclature
Truss Types
Features
Truss Inspection Checklist
Commonly Found Truss Defects
Section 10 – Pin-Connected Pratt Trusses
General
Features
Nomenclature
Pratt Truss Deterioration – Scenario
Pratt Truss Inspection Checklist
10 Movable Bridges
Section 1 – Inspection
Introduction
Inspection
Section 2 – Operator’s House
General
Section 3 – Swing Bridges
General
Operational Sequence
Special Inspection Items
Common Problems
Section 4 – Bascule Bridges
General
Operational Sequence
Special Inspection Items
Common Problems
Section 5 – Vertical Lift Bridges
General
Operational Sequence
Special Inspection Items
Common Problems
Section 6 – Electrical Inspection
General
Power Supply
Control Systems
Remote Control
Section 7 – Signal System
General
Section 8 – Mechanical Inspection
General
Hydraulic Operations
Section 9 – Lubrication
General
Items to Inspect
Section 10 – Fenders and Dolphins
General
11 Tunnel Inspections
General
Safety
Tunnel Inspections
Tunnel Inspection Checklist
12 Culvert Inspections
General
Safety
Inspection
Material Specific Inspection Requirements
Box/Arch Culverts
Timber and Rail-topped Culverts
Attachments
Conclusion
Culvert Inspection Checklist
13 Emergency Inspections
General
Fire
Flood
Derailments
Impact Damage
Catastrophic Events
Emergency Bridge Inspection Checklist
14 Post-Earthquake Inspections
General
Track and Roadbed
Drainage
Steel Bridges
Concrete Bridges and Substructures
Timber Trestles
Movable Spans
Culverts
Retaining Walls
Tunnels
Other Structures
Glossary
CHAPTER 1

BRIDGE INSPECTION PRACTICES

General

The Bridge Inspector is the person employed by the railroad who has the primary and important
responsibility to inspect structures, to determine the conditions of those structures, and to detect
conditions that impact the ability of those structures to safely carry the loads and forces imposed
on them by train and other loads. The Bridge Inspector is the first line of defense between the
effects of age and deterioration and a potentially catastrophic structural failure.

It is vitally important that the Bridge Inspector act with knowledge, judgment, and dependability
in the inspection and reporting process. It is, therefore, mandatory that the Bridge Inspector
carefully and diligently follow all inspection procedures for each and every bridge inspection.

Criteria for the limits of acceptable/tolerable conditions must be provided by


the Railroad owner, or his authorized representative, based on the principles
of acceptable rating contained in the appropriate chapter of the American
Railway Engineering & Maintenance-of-Way Association’s (AREMA)
Manual for Railway Engineering.

Safety Practices

Bridge inspection is inherently dangerous and requires continual attention to safety by each
member of the inspection team. The single most important consideration for inspecting bridges
safely is the individual inspector’s concern for creating a safe working environment through
attitude, alertness, common sense, planning, and training. Good work habits which lead to a safe
working environment include:

ƒ Keeping well rested and alert.


ƒ Maintaining health and physical conditioning.
ƒ Using proper tools and lighting.
ƒ Keeping work areas neat and uncluttered.
ƒ Establishing systematic procedures concerning what to expect of one another.
ƒ Following safety rules and regulations established by each individual railroad.
ƒ Safety rules and regulations in the U.S. established by the Occupational Safety and
Health Administration (OSHA) in 29 CFR, Chapter XVII, Article 1926 and the Federal
Railroad Administration (FRA) Title 49 CFR, Part 214, Railroad Workplace Safety.
ƒ For Federal Operated Lines in Canada, safety rules and regulations established by
Canada Labour Code (CLC), including CLC Part II, Sec. 12.10 (1) thru 12.10 (5) for fall
protection.
ƒ For short line railroads in Canada that operate exclusively in one province, safety rules
and regulations established by the Provincial Workman’s Compensation Board.
ƒ Using common sense and good judgment.
ƒ Avoiding alcohol and drugs.

The bridge inspectors are ultimately responsible for their own safety. The bridge inspector's
responsibilities also include:

ƒ Recognition of physical limitations.


ƒ Knowledge of rules and requirements of job.
ƒ Safety of fellow workers.
ƒ Reporting an accident immediately (or in some cases within 24 hours).

Personal Protection

It is important to dress properly for the job. Field clothes should be properly sized for the
individual, and they should be appropriate for the climate. For general inspection activities, the
inspector must wear approved safety shoes. For climbing bridge components, the inspector
should wear boots with a steel shank (with non-slip soles without heavy lugs), as well as leather
gloves. Wearing a tool pouch enables the inspector to carry tools and notes with hands free for
climbing and other inspection activities.

While safety equipment is designed to prevent injury, the inspector must use this equipment for
his or her protection. Some common pieces of safety equipment are:

ƒ Hard hat - provides protection from falling objects and protects the inspector's head from
accidental impact with bridge components.
ƒ Reflective safety vest - essential when working near traffic conditions.
ƒ Safety goggles - eye protection is necessary when the inspector is exposed to flying
particles; glasses with shatterproof lenses are not adequate since side protection is not
provided; extra care should be taken when climbing with bifocals or blended lenses.
ƒ Fall protection or fall restraint equipment.
ƒ Life jacket - should always be worn when working over water (if not tied off) or in a
boat.
ƒ Dust mask - protects the lungs from dust.
ƒ Respirator - protects the inspector from harmful airborne particulates and contaminants
from sand blasting, painting, and exposure to dust from pigeon droppings.
ƒ When Fall Protection equipment is used, a rescue plan must be in place, although such
equipment is exempt when FRA Regulation 49 CFR 214 Bridge Workers Safety Rule
214.103(b) (2) is followed.
ƒ Gloves - protect the hands.

Accident Prevention

The two major causes of accidents are human error and equipment failure. Human error can be
reduced by acknowledging that we all make mistakes and planning ahead to minimize their
effects. Equipment failure can be reduced by providing inspection, maintenance, and update of
equipment.

Be aware of the following causes of accidents:

ƒ Improper attitude - distraction, carelessness, and worry over personal matters.


ƒ Personal limitations - lack of knowledge or skill; exceeding physical capabilities.
ƒ Physical impairment - previous injury, illness, side effects of medication, alcohol, or
drugs.
ƒ Boredom or distractions - falling into an inattentive state while performing repetitive,
routine tasks.
ƒ Thoughtlessness - lack of safety awareness and not recognizing hazards.
ƒ Short-cuts - sacrificing safety for the sake of time.
ƒ Faulty equipment - damaged ladder rungs, worn ropes, and frayed cables.
ƒ Improper or loosely fitting attire.

Safety Precautions

Some general guidelines for safe inspections include:

ƒ Avoid use of intoxicants or drugs - impairs judgment, reflexes, and coordination.


ƒ Medication - prescription and over-the-counter medications can cause unwanted and
dangerous side effects including drowsiness or dizziness.
ƒ Electricity - all cables and wires should be assumed to be live; all power lines should be
de-energized and grounded or protected.
ƒ Assistance - always work in pairs.
ƒ Inspection over water - a safety boat, equipped with a life ring and radio communication,
must be provided when working over bodies of water if alternative safety equipment is
not provided. (scaffolds, fall protection equipment, etc.)
ƒ Waders - caution should be used when wearing waders since they can fill with water,
making swimming impossible.
ƒ Inspection over traffic - if working above traffic or waterways cannot be avoided or when
working on open deck structures, then tools and notebooks should be tied off.
ƒ Traffic control may be required for inspection of railroad bridges over roadways or
navigable waterways.
ƒ Entering dark areas - breathing the dust of pigeon droppings may be a health risk.
Always have adequate lighting to ensure personal and physical safety.

Climbing Safety

There are three basic areas of preparation necessary for a safe climbing inspection. The first
basic area is the organization of the inspection:

ƒ Climbing strategy - climbing time should be minimized.


ƒ Inspection plan - the inspector should know exactly where to go, what will need to be
done, and what tools will be needed.
ƒ Weather conditions – rain, snow, ice, or high winds may warrant postponement of bridge
inspections.
ƒ Traffic - should not be obstructed.

Second, inspection equipment should be checked for proper use and condition:

ƒ Ladders - accidents involving ladders are the most common, check ladder positioning and
condition.
ƒ Scaffolding - should be checked for height, load capacity, cracks, loose connections, and
weak areas, do not cantilever scaffolding.
ƒ Timber planks - two or more planks securely cleated together should be used; plank ends
should be securely attached to their supports.
ƒ Inspection vehicles - platform trucks, bucket trucks, and snooper type vehicles should be
used if possible.
ƒ Catwalks and travelers - permanent inspection access devices should be used when
available.
ƒ Rigging - the inspector should be familiar with proper rigging techniques and should not
have "blind" trust in the riggers.

Third, the inspector must be mentally prepared for a climbing inspection. A good safety attitude
is of foremost importance. Three precautions that must be addressed are:

ƒ Avoid emotional distress - do not climb when emotionally upset or when lacking self-
control.
ƒ Self-awareness - always know where you are and what you are doing when climbing.
ƒ Confidence - do not do anything you are not confident of doing safely, and do not hide
the fact that something was not inspected.

Confined Spaces

Inspection of box girder bridges, steel arch rings, arch ties, cellular concrete structures, and long
culverts often involves confined spaces. A thorough discussion of safety measures required when
working in confined spaces is presented in Chapter 2 of this Handbook.

Standard Tools

In order for the inspector to perform an accurate and comprehensive inspection, the proper tools
must be used. Standard tools that an inspector should have available at the bridge site can be
grouped into seven basic categories:

1. Tools for cleaning.


2. Tools for inspection.
3. Tools for visual aid.
4. Tools for measuring.
5. Tools for documentation.
6. Miscellaneous equipment.
7. Special equipment.

Tools for Cleaning:

Tools for cleaning should include:

ƒ Wisk broom - used for removing loose dirt and debris.


ƒ Wire brush - used for removing loose paint and corrosion from steel elements.
ƒ Scrapers (2 inch) - used for removing corrosion or growth from element surfaces.
ƒ Flat bladed screwdriver - used for general cleaning and probing.
ƒ Shovel - used for removing dirt and debris from bearing areas.

Tools for Inspection:

Tools for inspection should include:

ƒ Pocket knife (where permitted by railroad) - used for general duty.


ƒ Ice pick - used for surface examination of timber elements.
ƒ Increment borer or shell and void indicator - used for internal examination of timber
elements.
ƒ Chipping hammer with leather holder (geologist’s pick - used for loosening dirt and rust
scale, sound concrete, and checking for sheared or loose fasteners).
ƒ Plumb bob - used to measure vertical alignment of a superstructure or substructure
element.
ƒ Tool belt with tool pouch - used for convenient hold and access of small tools.

Tools for Visual Aid:

Tools for visual aid should include:

ƒ Binoculars - used to preview areas prior to inspection activity and for examination at
distances.
ƒ Flashlight - used for the examination of dark areas.
ƒ Lighted magnifying glass (e.g., five times and ten times) - used for close examination of
cracks and areas prone to cracking.
ƒ Inspection mirrors - used for inspection of inaccessible areas (e.g., underside of deck
joints).
ƒ Dye penetrant - used for identifying cracks and their lengths.
ƒ Paint remover.
Tools for Measuring:

Tools for measuring should include:

ƒ Pocket tape (6 foot rule) - used to measure defects and elements and joint dimensions.
ƒ 100 foot tape - used for measuring component dimensions.
ƒ Calipers - used for measuring the thickness of an element beyond an exposed edge.
ƒ Optical crack gauge - used for precise measurement of crack widths.
ƒ Paint film gauge - used for checking paint thickness.
ƒ Tilt meter and protractor - used for determining tilting substructures and measuring the
angles of bearing tilt.
ƒ Thermometer - used for measuring ambient air temperature and superstructure
temperature.
ƒ 4 foot carpenters level - used for measuring level and plumb.

Tools for Documentation:

Tools for documentation should include:

ƒ Inspection forms, clipboard, and pencil - used for record keeping for an average bridge.
ƒ Field books - used for additional record keeping for complex structures.
ƒ Straight edge - used for drawing concise sketches.
ƒ Digital or 35 mm camera - used for visual documentation of the bridge site and
conditions.
ƒ Chalk or markers - used for element and defect identifications for improved organization
and photo documentation.
ƒ Center punch - used for applying reference marks to steel elements for movement
documentations (e.g., bearing tilt and joint openings).
ƒ “P-K” nails - Parker Kalon masonry survey nails used for establishing a reference point
necessary for movement documentation of substructures and large cracks.

Miscellaneous Equipment:

Miscellaneous equipment should include:

ƒ “C”-clamps - used to provide a “third hand” when taking difficult measurements.


ƒ Penetrating oil - aids removal of fasteners, lock nuts, and pin caps when necessary.
ƒ Insect repellant - reduces attack by mosquitoes, ticks, and chiggers.
ƒ Wasp and hornet killer - used to eliminate nests and hives to permit inspection.
ƒ First-aid kit - used for small cuts, snake bites, and bee stings.
ƒ Toilet paper - Used for “emergencies” (better safe than sorry).
ƒ Water – to ensure proper hydration.
Special Equipment:

For the routine inspection of an average bridge, special equipment is usually not necessary.
However, with some structures, special inspection activities require special tools. These special
activities may be subcontracted by the agency responsible for the bridge. The inspector should
be familiar with special equipment and applications.

ƒ Survey Equipment
Special circumstances may require the use of a transit, a level, an incremental rod, or
other survey equipment. This equipment establishes a component’s exact location
relative to other components, as well as established reference points.

ƒ Nondestructive Testing Equipment


Nondestructive testing (NDT) is the in-situ examination of materials for structural
integrity without damaging the materials. NDT equipment allows the inspector to “see”
inside a bridge element and assess deficiencies that may not be visible with the naked
eye. Generally a trained technician is necessary to conduct NDT and interpret the
results.

ƒ Underwater Inspection Equipment


Underwater inspection is the examination of substructure units and the channel below
the water line. When the waterway is shallow, underwater inspection can be performed
with a simple probe. Probing can be performed using a piece of reinforcing steel, a
survey rod, a folding rule or even a tree limb.

When the waterway is deep, underwater inspection must be performed by trained divers.
This requires special diving equipment and may include other equipment such as a
working platform, fathometer, ground penetrating radar, air supply systems, radio
communication, and sounding equipment.

Other Special Equipment:

An inspection may require special equipment to prepare the bridge prior to the inspection. Such
special equipment includes:

ƒ Air/water jet equipment – used to clean surfaces of dirt and debris.


ƒ Sand or shot blasting equipment – used to clean steel surfaces to bare metal.
ƒ Drilling and grinding equipment.

Methods of Access

The two primary methods of gaining access to difficult to reach areas of a bridge are access
equipment and access vehicles. Common access equipment includes ladders, rigging, and
scaffolds. Common access vehicles include man lifts, bucket trucks, and snooper type vehicles.
In most cases, using a man lift or bucket truck will be less time consuming than using a ladder or
rigging to inspect a structure. The time saved, however, must offset the high cost associated with
operating access vehicles and the need for track time.

Access Equipment

The purpose of access equipment is to position the inspector close enough to the bridge
components so that a “hands-on” inspection can be performed. The following are some of the
most common forms of access equipment.

Ladders

Ladders can be used for inspecting the underside of a bridge or for inspecting substructure units.
However, a ladder should be used only for those portions of the bridges that can be reached
comfortably, without undue leaning.

Rigging

Rigging of a structure consists of cables and platforms. Rigging is used to gain access to floor
systems and the bottom of main load-carrying members in areas where access by other means is
not feasible or where special inspection procedures are required (e.g., nondestructive testing and
pin removal). Rigging is often used over water, over busy highways or railroads where sufficient
clearance exists, and for bridges that are over 40 feet high.

Scaffolds

Scaffolds are generally more mobile than rigging. They provide an efficient access alternative
for structures that are less than about 40 feet high and over level ground with little or no traffic.

Boats or Barges

A boat or barge may be needed for structures over water. A boat can be used for some
inspection, as well as for taking photographs. A barge can be used as a work platform for
underwater inspection, or to support ladders, scaffolding, etc.

Climbers (Powered Personnel Baskets)

Climbers are mobile inspection platforms that “climb” steel cables. They are well suited for the
inspection of high piers and other long vertical faces of bridge members. Climbers are sometimes
referred to as “spiders”.

Floats

A float is a wood plank work platform hung by ropes. Floats are generally used for access in
situations where the inspector will be at a particular location for a relatively long period of time.
Bosun (or Boatswain) Chairs

Bosun chairs are suspended with a rope and can carry only one inspector at a time. They
can be raised and lowered with block and tackle devices.

Climbing

On some structures, if other methods of access are not practical, inspectors must climb the bridge
elements. Safety awareness should be foremost in the inspector’s mind when utilizing this
technique. Climbing can be divided into two categories. The first category is free climbing, in
which the inspector climbs freely, unsecured to the bridge (where permitted by regulation and
the railroad company). The second category employs rappelling techniques and safety
equipment.

Access Vehicles

There are also many types of vehicles available to assist the inspector in gaining access to bridge
elements. The following are some of the most common types of access vehicles.

Man Lifts

A man lift is a vehicle with a platform or bucket capable of holding one or more inspectors. The
bucket is attached to a hydraulic boom that is mounted on a carriage. An inspector “drives” the
carriage using controls in the bucket. This type of vehicle is usually not licensed for use on
highway. However, some man lifts are nimble and can operate on a variety of terrains.

Bucket Trucks

A bucket truck is similar to a man lift. However, a bucket track can be driven on a highway, and
the inspector controls only the bucket. Outriggers are sometimes extended from the chassis of
the vehicles to help maintain stability, allowing greater reach and turning range. Some bucket
trucks can move along the bridge during inspection activities. Most bucket trucks also have
multiple booms with some models providing reach of up to 60 feet.

Snooper Type Vehicles

A snooper is a specialized bucket truck with an articulated boom designed to reach under a
structure while parked on the deck. A rotating turret provides maximum flexibility, and
outriggers with wheels allow the truck to be moved during operations. Usually the third boom
has a capacity for extending and retracting, allowing for greater reach under a structure.
Emergency

If the inspector discovers a bridge condition that affects the integrity of the
bridge under train loads, contact the railroad dispatcher and responsible
authority to stop any trains and arrange for immediate repairs.
CHAPTER 2

CONFINED SPACES

General

Inspection of box girder bridges, steel arch rings, arch ties, cellular concrete structures, and long
culverts often involves confined spaces. Strong consideration should be given to establishing, by
policy, that "ALL" culverts, regardless of size, length, or design are confined spaces. Inspectors
should observe all State, Federal, and Operating Railroad safety requirements.

Safety

There are three major safety concerns when inspecting a confined space:

ƒ Lack of oxygen - oxygen content must remain above 19.5% for the inspector to
remain conscious.
ƒ Toxic gases - generally produced by work processes such as painting, burning, and
welding.
ƒ Explosive gases - materials such as natural gas and methane are produced by the
natural oxidation of organic matter.

When a confined area must be inspected, some basic safety precautions should be followed:

ƒ Test oxygen and other gases at 15 minute intervals.


ƒ Avoid use of flammable liquids.
ƒ Position inspection vehicles away from the area to avoid carbon monoxide fumes.
ƒ Operations involving gasoline or toxic gases should be performed "down wind" of the
operator and the inspection team.
ƒ Use approved air-breathing apparatus when ventilation is not possible or when
detection equipment is not available.
ƒ Adequate lighting and lifelines are required when entering culverts.
ƒ Inspection should be performed in pairs, with a third inspector remaining outside of
dark or confined areas.
ƒ Have communications between personnel in the space and personnel outside the
culvert.

While having a comprehensive permit program in place makes it unlikely that a confined space
rescue will be required, it is still important to plan for the unexpected. Items to consider when
planning for a confined space rescue:

ƒ Evaluate the confined space.


o Internal configuration – open or obstructed.
o Elevation – elevated or at grade.
o Portal size – Restricted or unrestricted.
o Space access – Horizontal or vertical.
ƒ Required response time.
ƒ If rescue team is not on site, time it will it take to get there.
ƒ Rescue team availability during all required hours.
ƒ Rescue team meets all applicable requirements.
ƒ Adequate communications in place to transmit a rescue request without delay.
ƒ Where airline respirators are required, have procedures in place including an ample
supply of air cylinders.
ƒ If the space has vertical entry over five feet in depth, the rescuers need to be properly
trained in climbing, rope work, or elevated rescue.
ƒ Necessary skills on-site in medical evaluation, patient packaging, and emergency
response.
ƒ Necessary equipment required to perform a rescue, and who provides.

Culverts

A culvert is any structure or conduit designed exclusively for the direction or passage of fluids
along or through the right of way. By establishing that culverts are confined spaces, it serves to
warn the entrant of possible dangers due to:

ƒ Design (single entry or exit).


ƒ Construction (curved, smooth sides, narrow, uneven/holes in floor surface not
obvious because of water flow).
ƒ Location (isolated area).
ƒ Atmosphere (low oxygen because of high corrosion, presence of toxic gases).
ƒ Materials (debris buildup within, water buildup outside).
ƒ Other conditions (structure deterioration, i.e.: loose rock/masonry, root systems, worn
out floor, metal protrusions account failure).

Having ascertained (due to training in confined space entry) that none of the aforementioned
hazards exist, the culvert may be entered without a confined space permit.

Tunnels

Tunnels used for vehicle traffic, or train operations, are not considered a confined space, but
more so an enclosed space. However, conditions such as derailments, work programs, ice
conditions, could result in a tunnel becoming a permit required confined space. Tunnels,
generally, are to be considered as enclosed spaces with work procedures.

Movable Bridges and Turntables

Movable bridges and turntables should, by policy, be considered a confined space.


ƒ They are not designed for human occupancy except for the purpose of doing work.
ƒ They have limited or restricted means of access/egress.
ƒ There are mechanical hazards.
As the mechanical hazard presents the greatest danger to personnel, full compliance of a LOCK
OUT/TAG OUT POLICY must be undertaken.

In addition, there are potential hazards which could include flammable/combustible atmosphere,
toxic atmosphere, insects and vermin.

"NOTE" weigh scales may have the same characteristics as movable bridges and turntables.

Other Structures

There may be other structures on the property that may be similar in design to culverts, but they
may not have to be considered as confined spaces. These would be structures such as pedestrian
underpasses and cattle passes.

ƒ Pedestrian underpasses, due to design for human occupancy, would have a sound
floor system, have sufficient lighting, ventilation if required, be free of any hazards.

ƒ Cattle passes, due to volume of animal traffic, would be clear of debris, have a sound
floor system, and in general be free of any hazards.

Emergency

If the inspector discovers any condition that affects the integrity of the
structure under train loads, contact the railroad dispatcher and responsible
authority to stop any trains and arrange for immediate repairs.
CHAPTER 3

SITE CONDITIONS

General

A number of items other than the condition of the substructure, superstructure, and track should
be observed and noted during a thorough bridge inspection. This section briefly discusses site
conditions that can impact a bridge and its approaches.

Utilities

Note the presence of utilities near the bridge or attachments to the bridge. Note the presence of
overhead utility lines and underground utility markers. Although unlikely, the potential exists
for unauthorized utility presence (encroachment) on railroad property.

Pipelines and other underground utilities can be exposed by streambed scour and degradation.
Look for marker signs.

Adjacent Property Development

For railroads over automobile-truck routes, development could render the railroad bridge length
inadequate to provide the needed lane width below the railroad, thus leading to eventual bridge
replacement. Increases in truck traffic could also render the vertical clearance inadequate.
Introduction of roadway salts as a result of development could negatively impact the railroad
bridge substructure and superstructure elements. For railroads over streams, development could
change the drainage basin characteristics (faster runoff) necessitating an increased stream
opening to prevent overtopping or washout. The inspector should note change of condition due
to any development in the inspection report.

Vegetation Growth

Vegetation growth encroaching on the track at bridge approaches can be to the detriment of
facility operation. Vegetation under the bridge and adjacent to the substructure elements can be
destructive. Tree limbs may become entangled around lateral bracing and may interfere with
access to superstructure elements by the bridge inspector. Larger trees near the bridge can
potentially drop limbs on the structure during wind or lightning storms, thus causing damage.
The inspector should point out conditions of vegetation in the inspection report with
recommendations that heavy vegetation (trees and vines) under and adjacent to the bridge be
cleared.

Drift accumulation can pose a number of risks to a structure including fire hazards, scour and
erosion, redirection of the channel, and possibly exerting enough pressure on the bridge during
high flows to push it out of alignment or cause structural damage.
Erosion

Embankment or streambed erosion related to scour at contractions and/or obstructions in the


waterway opening near, and particularly upstream, of bridge openings is a major cause of bridge
failures. The inspector should be aware of changes in the embankment and/or streambed that
could be detrimental to the stability of the bridge and note these conditions (changes) in the
inspection notes. Note conditions upstream and downstream of the bridge that could impact
scour such as:

ƒ Development of sandbars.
ƒ Drift (dead tree limbs) and debris build up at substructure elements.
ƒ Beaver dams.
ƒ Tree growth or vegetation growth that may be causing the stream channel to migrate
towards an abutment.
ƒ Note the existence of “head cuts” that appear as a vertical drop in the streambed and
resemble a small waterfall. Head cuts can move upstream quickly during a storm
undermining shallow foundations and eroding embankments at bridge ends.
ƒ Note approximate stream level and velocity at time of inspection.
ƒ During winter months note any build up of ice flows at the substructure elements.
ƒ Note any embankment slides, overall slope stability, exposed footings, etc.
ƒ Note the condition of riprap and/or other erosion control installations below the structure.

Adjacent or Parallel Bridge/Culvert Structures

Drift and debris build up at an adjacent bridge or culvert could cause an overtopping situation at
a parallel structure. Note the presence of a small ditch/stream parallel to track that may be
impacting the stability of the railroad structure approach embankment. Note the condition of
special access roads leading to the bridge.

Track/Bridge Geometry

Closely observe track and bridge geometry. Look for unusual changes in rail, girder, and
walkway railing alignment both in the horizontal and vertical planes.
These conditions could be an indication of a more serious streambed or embankment erosion
problem or structural condition.

Note if pier/bent protection work is required or whether cleaning and straightening of the channel
are necessary. Note whether the bent alignment obstructs or deflects normal flow and if
revetment or deflection dikes are needed. Observe if an island is forming on the downstream
side of the bridge, as this is an indication of possible scour problems.

ROW Fences

Be aware of railroad property lines. Note conditions where new fence has been installed or
existing fence has been removed. Note the condition of fences below the bridge that may be for
the purpose of livestock control. Certain fence types could restrict stream flow and collect
debris.

Vandalism

Note any vandalism observed. This may be graffiti, signage damage, presence of debris,
campfire remains, or other damage to bridge or nearby utilities. Note any trespassing that may
be observed during the course of the inspection.

SITE CONDITIONS CHECKLIST

Utilities

___ Underground
near the bridge
attached to the bridge
clearly marked
___ Overhead
condition
attachments to utility poles
___ Unauthorized utilities present

Adjacent Property

___ Recent development impacting on train operations or right-of-way


___ Changes in underlying roadways
overhead clearance changes
widening affecting bridge structure
___ Changes in drainage basin development

Vegetation Growth

___ Encroachment on bridge approaches


___ Beneath bridge structure

Erosion

___ Scour
Upstream
Downstream
Sandbar development
Head cuts
Foundations exposed or undercut
___Adjacent slopes
Embankment slides
Riprap conditions

Adjacent or Parallel Bridge/Culvert Structures

___ Debris or drift buildup that could cause overtopping an erosion at structure
___ Erosion of adjacent ditch or stream endangering railroad embankment
___ Access roads adjacent to or close to railroad structure

Track/Bridge Geometry

___ Structure alignment


Vertical
Horizontal

ROW Fences

___ Railroad property lines


Condition of existing fences and whether any have been removed
Note any new fences, locations, infringement on railroad ROW
___ Fences
Conditions for livestock control
Buildup of debris

EMERGENCY

If the inspector discovers a site condition that affects train operation or safety, contact the
railroad dispatcher and responsible authority to stop any trains and arrange for immediate
repairs.
CHAPTER 4

BRIDGE LOADS AND FORCES

General

The bridge inspector should know and understand the different types of loads that a structure, a
part, or a single member of a bridge structure may be required to carry. The bridge inspector
must also understand how those loads are transmitted through the structure as types of forces
(axial and/or bending, tension and compression, shear, torsion, etc.).

Types of Loads and Forces

The basic loads used to design and evaluate bridges are as follows:

ƒ Dead Load – Load due to the weight of the structure itself (includes all bridge
components plus track, ballast, walkway, and any other permanent attachments).
ƒ Live Load – Weight of the portion of the train on the bridge. Current railroad design
loads are for 80,000# axles, except for certain steel bridge members which are designed
for 100,000# axles. Design live loads may be different on an individual railroad.
ƒ Impact Load – An additional load placed on the structure to account for the dynamic
effect of a train moving over a bridge. It is a percentage of the Live Load, and increases
with train speed. Impact is not used in the design of timber structures but should be
considered as a factor in evaluations.
ƒ Centrifugal Force – A horizontal force on structures with curved track, applied toward the
outside of the curve to account for the effects of the train as it moves around the curve.
Force increases with train speed and the degree of curvature.
ƒ Wind Force – Wind force acting horizontally on the bridge superstructure and support
elements, and on the train itself.
ƒ Longitudinal Force – A force due to the tractive effort or braking of the train.
ƒ Seismic Forces – Forces due to earthquakes.
ƒ Stream Flow Forces – Forces due to the force water may exert on the side of a bridge
structure or its foundation. Forces will change due to height, velocity of flow, debris, and
drift.
ƒ Ice Forces – Horizontal forces exerted on a structure due to ice pack or ice flows.
ƒ Forces from Continuous Welded Rail – Forces generated due to the restraint of rail
expansion or contraction.
ƒ Lateral Earth Force – Horizontal force exerted on the substructure by the surrounding
earth.
ƒ Hydrostatic Force – A force generated by a differential elevation of body of water or
ground water.

Combinations of these loads are applied to a bridge to determine the design forces. Individual
bridge members are then sized accordingly as part of the design process. Changes over the years
in track geometry, site conditions, etc., can change the loads applied to the structure and its
individual components, and should be noted during inspections. Also note changes from previous
inspections.

During the design process, once all of the appropriate loads have been identified, they are
applied to the structure, and the forces that are generated are used to design all of the bridge
components. It is helpful for the inspector to have a basic knowledge of the different types of
forces acting on a bridge and its components so that he/she may better understand the relative
importance of both the component being inspected as well as any defects that may be identified.
The three most common types of forces that are found in railroad bridges are as follows:

ƒ Tension Force – An axial or bending force that tends to stretch or lengthen a member.
Figure 4-1

ƒ Compression Force – An axial or bending force that tends to shorten or compress a


member. Figure 4-2

ƒ Shear Force – A force which causes one section of a member to want to slice past another
section in a sort of scissors action. Figure 4-3

Different members of a bridge may have one or more of the above forces acting upon it
depending on the type of structure, and how the loads are applied. When bridge spans are loaded
with a train, they deflect or bend downward at or near the center of the span. The deflection may
not be visible to the eye, but it is happening. The following two diagrams show exaggerated
deflection due to a train, of a simple span and a continuous span (in a continuous span, each span
does not end at a pier but is continuous over the pier). Figures 4-4 & 4-5

This bending action creates forces or stresses in the structure as it resists the loads imposed by
the train. For a simple span, the top portion of the bridge goes into compression (the material is
being pushed together) and the bottom goes into tension (the material is being stretched apart).
The ability of the material, whether it is steel, concrete, or timber, to resist the compression and
tension, helps determine the strength of the bridge.

In addition to the bending, there are also shear forces at work when the bridge is loaded.
Remember the diagram showing shear force. The bearings at the piers and abutments are
holding the bridge up and the train is trying to push it down. This is creating a shear force where
the bridge must be strong enough to resist the downward force of the train.

One other force that has not yet been mentioned is torsion. Torsion is a twisting force as shown
in the following diagram. Figure 4-6

Torsion in railroad bridges is usually limited to machinery on moveable bridges and turntables.
It is easy to envision torsion forces in a drive shaft that rotates to operate the bridge. Torsion
forces will not be addressed any further in this section.

To better understand these forces, examine the following diagram. Figure 4-7
There is a large shear force between each bearing and the train axles as indicated in the shear
diagram. This is the scissor type force created when the train axles push down on the span and
the bearings hold the span up. Shear is greatest near the ends of the span. Shear forces are
carried by the webs on girder and beam spans and by the diagonals and posts on truss spans.

Now refer to the moment diagram to consider the bending. As the span is loaded, it bends more
in the middle, especially as individual axles get close to mid-span. In this case you can see that
the bending (moment) is greatest near the middle of the span as common sense would indicate.
The important thing to remember here is that bending (moment) is the greatest near mid-span
and decreases to almost nothing adjacent to the bearings on simpler spans. Remember that on
simple spans, the bending creates tensile forces in the bottom of the span and compressive forces
in the top. In Figure 4-8 bending moment forces are carried by the top and bottom flanges on
girder and beam spans and by the top and bottom chords on truss spans.

It is important to understand how a bridge reacts to these forces to better evaluate whether
defects you find on inspections may be critical. In order for the bridge to remain standing, the
size, shape, and type of material it is made of must be adequate to resist failure in tension,
compression, shear, or bending (there are also factors of safety involved).

Fatigue

Fatigue is the effect on structures of being loaded and unloaded many times. The two major
factors affecting fatigue life with respect to loading are: the magnitude of the stresses caused by
the loading (which increase proportionally with the weight of the train), and the number of cycles
(which is dependent on the type and number of trains as well as the length and orientation of an
individual bridge component). Some bridge members that are extremely susceptible to fatigue
are short girder spans or floor system panels, and connections. Figures 4-9 & 4-10

Trusses also have members that are very susceptible to fatigue due to cyclical loading. Figure 4-
11 shows what happens to individual truss members as a train moves across the bridge.
Members that are in tension are shown in yellow and compression members are shown in blue.
Members with no train load reaction are left uncolored. Note that the bottom chord stays in
tension during the entire passage of the car and the top chord stays in compression. Several of
the vertical and diagonal members however cycle between tension and no load or tension and
compression. These members are subject to potential fatigue damage depending on the
magnitude of the train load induced stresses. Figure 4-11

Other factors that affect fatigue life are the type and quality of the materials, temperature and the
environment, structural details (sudden change in cross-section or shape, etc.), and connections
(welds, rivets, bolts). During an inspection, it is important to note any conditions that may be
fatigue related. For more information on fatigue, see the Steel Bridges section of this Handbook.
Load Paths

Understanding load paths helps the inspector to understand how loads are being handled by each
component of a structure. This knowledge is valuable to understanding where problem areas
might occur, as well as their relative importance. As an example, consider a through plate girder
span with on open timber deck. See AREMA Manual for Railway Engineering, Figure 15-7-2.
The load of the train axles is transmitted through the rail to the cross ties which rest on the
stringers. That load is then transmitted to the floor beams through connections at the ends of the
stringers. The floor beams then carry the loads into the girders which finally carry the loads
directly to the bridge bearings and substructure.

Load Redistribution

Many structures have the ability or the tendency to redistribute the loads among different
members than were originally designed to carry the loads. This is known as “redundancy”. Look
at how any defect relates to the rest of the structure to help determine the cause.

Deflection

Deflection is a measurement of how far a beam or girder bends down when loaded with a train,
which is normal behavior. If movement seems excessive during the passage of a train, look
closely at the bearing areas and substructure to determine the nature of the movement. If it is
indeed deflection and it seems excessive, measure the deflection or make a note of it so it can be
further evaluated.

Bridge Ratings

Bridge Rating is a determination of the load, or loads, a bridge has the capacity to support in its
current condition. The bridge rating takes into account specific conditions found in the field such
as damage, corrosion, misalignment, etc. as well as the actual train traffic and speeds that the
bridge is supporting. Bridge rating should not be confused with “bridge design” which utilizes
standardized axle loads and spacing and the assumption that everything will be constructed
exactly as shown on the plans for the construction. If a bridge rating is to be performed, care
should be taken during the inspection to identify any conditions that may have changed since
construction or previous inspections, so that enough information is contained in the report to
perform an accurate load rating if desired.

Other Factors

Due to conditions found in the field, bridges can be subjected to loads not anticipated or
accounted for in the original design. Some of these loads can be caused by flat wheels, bad
joints or flat spots in the rail(s). Flat wheels obviously cannot be identified during inspection,
and since they will seldom impact at the same location, it is unlikely they will have much of an
adverse impact on the structure unless the defect is quite large. Impacts due to track defects,
however, can be a cause of major concern because they are always at the same location and will
generate an impact load with each passing wheel. This problem can have an even worse effect
if the bridge member being subjected to the impact load is susceptible to fatigue. Always
identify the location and nature of track defects or bad joints, including movable bridge joints
that could cause undue impacts to the bridge.
Figure 4-1 Tension Force
Figure 4-2 Compression Force
Figure 4-3 Shear Force
Figure 4-4 Simple Span Deflection
Figure 4-5 Continuous Deflection
Figure 4-6 Torsion
Figure 4-7 Shear and Moment Diagram
Figure 4-8 Bending Forces
Figure 4-9 Fatigue Cycles Short Span
Figure 4-10 Fatigue Cycles Longer Span
Figure 4-11 Fatigue Cycles on Truss Span
CHAPTER 5

BRIDGE NOMENCLATURE

General

To enable clear communication and understanding between bridge inspectors and those
performing maintenance and repairs, it is important that uniform terminology and methods for
identifying and locating the various components of a structure be used. It is particularly
important when discussing repairs, materials, defects, or emergencies by telephone to be able to
give a clear, concise, and correct description that will avoid confusion or misunderstandings.

Bridges are the broadest and most widely varied category of structures encountered on railroad
systems. In addition to railroad bridges, there are highway bridges, grade separations, signal
bridges, portions of track scales, turntables, and unloading structures that would fall within the
broad category of bridges. While differing in size and details of design, bridges of the same
basic type and construction materials will share similar component identification terminology,
regardless of the intended use of the bridge. This section will present the types of bridges
commonly found on railroad systems, along with diagrams to provide for clarity and future
reference.

Fundamentally, a bridge is a structure that provides a means to cross or span an obstruction. The
nature of the obstruction can be a waterway, a land feature such as a valley or ravine, another
railroad or a highway, or any other obstacle. The essential parts of a bridge are:

• Substructure. The abutments, piers, or other structures built to support the spans of a
bridge.

• Superstructure. The entire portion of a bridge structure which primarily receives and
supports rail, highway, or other traffic loads or facilities, and in turn transfers these loads
to the bridge substructure. The superstructure may consist of beam, girder, truss, trestle,
and other types of construction. (See Figure 5-1)

• Deck. The track, roadway, and attachments and incidental parts designed to directly
support and transmit traffic and facility loads to the superstructure.

Substructure Components and Terminology

Foundation

The foundation is the portion of the structure which transmits the entire load of both the
superstructure and substructure to the underlying soils. Depending on soil conditions at any
particular location, the foundation may consist of spread footings or deep foundations.
Spread footings are typically utilized where rock or hard soils are found at a relatively shallow
elevation.

Deep foundations may consist of driven piles, drilled shafts, caissons, or variations of same.
Deep foundations develop their load carrying capacity either by end bearing (transmitting the
load to a hard layer at the bottom of the element), side friction along the entire length of the
element, or a combination of these.

Abutment

An abutment is a substructure unit composed of stone, concrete, brick, or timber that supports the
end of a single span, or the extreme end of a multi-span superstructure, and in general retains or
supports the approach embankment. (See Figure 5-2) The following are the components of a
typical abutment:

• Bridge Seat. The top horizontal surface upon which the superstructure is placed and
supported.

• Backwall. The topmost vertical portion of an abutment above the bridge seat,
functioning primarily as a retaining wall for the approach embankment. It may also serve
as a support for one or more track ties at the end of the bridge.

• Wingwall. A retaining wall extension of an abutment that is intended to restrain and


hold in place the side slope material of the approach embankment. It may also serve to
deflect stream water and floating debris into the waterway and prevent embankment
erosion.

• Breast Wall or Stem. The portion of the abutment between the wings and beneath the
bridge seat that transmits the bridge loads to the foundation.

• Footing. The enlarged lower portion at the bottom of the abutment. In soft or unstable
soil conditions, piles may be driven, or other deep foundations utilized, to support the
footing.

Pier

A pier is a substructure unit composed of stone, concrete, brick, steel, or timber and is built in
shaft or block-like form to support the intermediate ends of the spans of a multi-span structure.
Examples of bridge piers are shown in Figure 5-3. The following are the components of piers
that would generally be encountered on railroads:

• Bridge Seat. The top horizontal surface upon which the superstructure is placed and
supported.

• Cap. The topmost horizontal portion of the pier. On a solid shaft pier, the cap, if
present, is of slightly larger horizontal dimensions than the balance of the pier shaft.
• Pier Wall or Stem. The portion of the pier between the footing and the pier cap.

• Footing. The enlarged lower portion at the bottom of the pier. In soft or unstable soil
conditions, piles may be driven, or other deep foundations utilized, to support the footing.

Bent

A bent is a supporting unit of a trestle or a viaduct-type structure made up of two or more


column or column-like (post) members connected at their topmost ends by a cap, strut, or other
member that holds them in their correct position. When piles are used as the column element,
the entire construction is designated as a “pile bent”. Alternatively, when the column elements
are constructed of sized timbers supported by a sill, the assemblage is termed a “framed bent”.
When fabricated from steel shape, the assemblage is termed a “steel bent”. Examples of bents
are shown in Figure 5-3. Important elements of bents are as follows:

• Cap. The topmost horizontal member serving to distribute the loads upon the columns.

• Pile Cap. The topmost horizontal member of a pile bent serving to distribute the loads
upon the piles and to hold them in their proper relative positions.

• Column, Post, Pile. An element situated in a vertical or nearly vertical manner,


generally having considerable length in comparison to its transverse dimensions.

• Sill. A base piece or member of a bent serving to distribute the column or post loads to
the foundation or mud sills.

• Mud Sill. A piece of timber, or unit composed of two or more timbers placed upon a soil
foundation as a support for a framed bent, or other similar member of a structure.

• Transverse or Sway Bracing. Members connecting the columns, sill, and/or caps, in
order to give rigidity to the complete assemblage in the plane located transverse to the
bridge alignment.

• Sash Bracing. Horizontal bracing struts between sway bracing panels located transverse
to the bridge alignment.

Tower

A tower is a four-sided substructure framework (two bents braced together longitudinally) in


a viaduct-type structure that supports the ends of two adjacent spans, or one complete span
(i.e., tower span) and the ends of two adjacent spans. The column members are braced and
strutted in tiers, and the planes of either two or four sides may be battered. This term may
also be used to designate the end supports of suspension spans, vertical lift spans, etc.
Superstructure Types, Components, and Terminology

Slab Bridge

This is the simplest type of superstructure, with the slab carrying the loads directly to the
abutments or piers. Construction of slabs varies from the natural stone roof used on the
earliest stone boxes to the reinforced and prestressed concrete designs of today. A typical
example of a slab bridge is illustrated in Figure 5-4.

Masonry Arch

The arch is another type of superstructure that was widely used to bridge larger streams,
roadways, and other obstructions in the early construction of railroads. The majority of these
were built of stone that was readily available at or near the construction site. Bricks were
also used, and in later years, cast in-place concrete arches were constructed. Many of the
stone arch structures built in the mid-1800’s are still in use today, carrying loads much
greater than originally anticipated. Examples are shown in Figure 5-5. Important elements
of a typical arch are included.

• Span. Horizontal distance between spring lines of abutments or piers.


• Rise. Vertical distance between the spring lines and crown.
• Spring Line. The inner edge of the surface or joint upon which the bottom end of the
arch rests.
• Arch Ring. The entire arch between spring lines.
• Crown. The highest part of the arch ring.
• Keystone. The highest wedge-shaped block or stone at the center of the arch ring.
• Spandrel. A wall or column resting on the arch ribs and supporting the deck.

Timber Trestle

Timber trestles are another type of structure that have been widely used because of availability of
material at or near the site when early railroad construction was at its peak. Many timber trestles
are still in service, having been renewed several times since originally constructed. Trestles are
constructed with either driven pile bents (Figure 5-6) or framed bents (Figure 5-7) as was briefly
discussed in the preceding section on substructures. Figure 5-8 illustrates one example of a
timber trestle (an open deck trestle) and the associated terminology.

Beam Span Bridge

These bridges consist of rolled steel I-shaped or H-shaped members, or reinforced concrete
member, two or more of which support the track or deck and carry the loads directly to the
abutments or piers. These structures may be open deck, where the ties rest directly on the top
flanges of the beams, or ballasted deck, where a concrete slab, steel plate or timber deck supports
the track and roadbed. Ballasted deck bridges utilize ballast retainers or parapet walls to hold the
ballast along the sides. Figure 5-9 shows a typical beam span bridge, and illustrates both open
and ballasted deck configurations.
Deck Plate Girder (DPG) Bridges

Deck plate girder bridges consist of two or more large beam-like members (i.e., girders)
fabricated by riveting, bolting, or welding plates and angles (or just plates) together. The deck
rests upon the top flange of the girders and may either be open (i.e., with the ties resting directly
on the girders), or solid. For solid decks, the deck may consist of a concrete slab, steel plate, or
timber deck that supports the track and roadbed, with ballast retainers or parapet walls to hold
the ballast along the sides. The load capacity of the girders is significantly increased through the
use of bearing and intermediate stiffeners, which serve to reinforce the web. These stiffeners
form rectangular “panels” in the web of the girder. The girders are tied together with bracing
that consists of top laterals, bottom laterals, and cross frames. The bracing is usually made up of
angles, channels and/or tee sections that are connected to gusset plates, which in turn are
attached to the girders at panel points. A typical deck plate girder bridge and its associated
terminology are illustrated in Figure 5-10.

Through Plate Girder (TPG) Bridges

Like deck plate girder bridges, through plate girder bridges consist of two or more large girders
fabricated by riveting, bolting, or welding plates and angles (or just plates) together. Web
stiffeners reinforce the girders and a system of lateral and cross braces tie the girders together
and provide stability. In this case, however, the track or deck is supported by a floor system that
is located below the top flange of the girders. Knee braces are used to provide support to the top
flange against buckling. The floor system itself is usually made up of floor beams and stringers.
The deck rests on the stringers that are framed into the floor beams. The floor beams are
attached to the girders at panel points, and transmit the track and deck load from the stringers to
the girders. The deck may be open (i.e., with ties resting directly on the floor system) or solid
(using a concrete slab, steel plate, or timber deck to support the track and ballast). Figure 5-11
shows a typical through plate girder bridge.

Truss Bridges

These bridges utilize trusses as the primary members of the superstructure. Trusses are open-
web frames that consist of jointed members so arranged that the frame is divided into a series of
triangular shapes. Due to its inherent strength and stability characteristics, the triangle is the
fundamental element in truss design. Figure 5-12 illustrates the various truss arrangements
typically used in railroad bridges. As this figure shows, a bridge may be either a through truss
(where the trains travel through the structure) or a deck truss (where trains travel over the top of
the structure).

The majority of truss railroad bridges are constructed of steel, but a few timber trusses still
remain in limited service for rail traffic. Timber trusses are also in use on some overhead
highway bridges, and generally have relatively low load limits. Therefore, the remainder of this
discussion of truss bridges will be primarily applicable to trusses constructed of steel.
The individual components of trusses may be solid rods, eye bars, pipe, tubing, rolled sections,
sections built up from plates and angles, or various combinations thereof. Rivets, bolts, welds,
pins, or a combination of these may be used to connect the truss components. Although the size
and shape of trusses will vary widely, many essential components will be common to all. The
majority of truss bridges carrying railroad traffic will be of the through truss type, divided
primarily between the riveted and pin-connected variety. Figure 5-13 illustrates the typical
component arrangement and terminology for through truss bridges.

Referring to Figure 5-13, the perimeter members of a truss consist of a top chord, bottom chord,
and end posts. The interior members of a truss that complete the triangular construction consist
of diagonals, intermediate posts, and hangers. These members are connected to gusset plates,
which form the panel points of the truss. The track is supported by a floor system, usually made
up of floor beams and stringers. The track rests on the stringers, and the stringers are framed into
the floor beams. The floor beams are attached to the trusses at panel points. The truss is
laterally braced by sway bracing, top laterals, and bottom laterals.

Movable Bridges

A movable bridge is a bridge of any type having one or more spans capable of being raised,
turned, lifted, or slid from its normal traffic service location to provide for the passage of
navigation.

Swing Bridge

A swing bridge is a span, usually of truss or plate girder construction, designed to be supported
solely on a pier at its center when its end supports have been withdrawn or released. It is
equipped to be turned in a horizontal plane once it is released from its end supports in order to
open the navigable waterway. When closed in the normal traffic position, the span is supported
at the center pier and at two outer rest piers or abutments. Swing bridges may be of center-
bearing or rim-bearing construction. See the Movable Bridge section of this Handbook for
further details. A typical swing bridge is shown in Figure 5-14. Components or mechanical
systems encountered on swing spans may include the following:

• Center Bearing. This is usually a bronze disc running in oil.

• Rack. Large toothed gear segments that are anchored to the center pier, concentric with
the center pivot, and are part of the mechanism by which the bridge is rotated.

• Pinion. A small, mechanically driven toothed gear that meshes with the rack and applies
the rotating force to the span through its shaft bearing attachments to the span.

• Balance Wheels. These run on a circular track on the outer edges of the center pier.

• Live Load Support Wedges. Wedge-shaped bearing blocks (usually mechanically


driven) that are placed under the outer ends of the bridge to lift and support the ends of
the span under traffic. Wedges may also be used under the truss or girder at the center
pier to remove all or part of the traffic load from the center bearing.

• Span Locks or Rail Locks. Mechanical devices that positively engage the swing span to
the fixed rest pier or approach rails when in the closed or normal traffic position.

Bascule Bridge

A bascule bridge is a span, usually of plate girder or truss construction, which lifts by rotating
vertically about a horizontal axle or trunnion or on a rolling surface. A counterweight is used to
offset the dead load of the leaf overhanging the trunnion, thus minimizing the power required to
open the bridge. A typical bascule bridge is shown in Figure 5-14.

Vertical Lift Span

This bridge consists of a movable span, usually of truss construction, with a fixed tower or
towers at each end. The span is connected to cables that pass over sheaves (pulleys) atop the
towers and connect to counterweights on the other side. The actual lifting is performed (usually
by electric motors) through the turning of the counterweight sheaves, or drums that wind
separate uphaul and downhaul cables. The general arrangement of a vertical lift span is
illustrated in Figure 5-14.

Turntables

Similar in principle to a swing bridge, turntables are usually found at engine houses or servicing
facilities. They are used to turn locomotives or other equipment, and to transfer them from one
to another of multiple tracks that radially extend away from the turntable. Turntables usually
consist of a plate girder span (either deck or through plate girder design) that is rotated on a
center bearing, with the ends supported by trucks (i.e., wheel assemblies) running on a circular
rail. Power for rotation is usually supplied by electric motors that drive one or more of the
wheels traveling over the circular rail. Turntables are usually placed in a concrete pit that has
circular walls and a sloped floor for drainage.

Transfer Tables

Transfer tables are found in shop or maintenance facilities and are used to move rolling stock or
track mounted equipment from one track to another. Transfer tables typically consist of a steel
beam or deck plate girder structure mounted on wheels and rails in a pit that allow the structure
to move transversly along the length of the pit and line-up the rails on the transfer table with the
approach rails on any track on either side of the pit. Drive mechanisms are similar to turntables.

Identifying Structures and Components

It is very important to establish a uniform system of identifying a structure as well as the


components that make up that structure. The system should be consistent from structure to
structure and should be known and understood by not only the inspectors, but also any other
individual who may need to read, interpret, or further evaluate the information contained in the
report. Following are some examples of items to be addressed in a standard identification:

• Bridge Number – Typically the bridge milepost, a unique number. Any sequential
numbering system can be used, but mileposts make the structure location easy to identify
and if a new structure is added, the new number is easily fit into the system.

• Abutment, Pier, & Span Numbers – These numbers typically increase in the direction of
ascending mileposts on the railroad. Starting at the lower milepost end of the bridge
would be Span 1, Span 2, Span 3…..and so on. In like fashion, Abutment 1, Pier 2, Pier
3, Pier 4…..Abutment 10 or Bent 1, Bent 2, Bent 3, etc.

• Component Numbers – Components numbered along the length of a bridge or span are
typically perpendicular to the track and numbered once again with increasing mileposts.
For instance, an intermediate floor beam of a span in the middle of a multi-span bridge
might be Floor Beam 5 of Span 3. Members that run parallel to the track such as
stringers or girders can be numbered from left to right, once again facing ascending
mileposts.

• Compass directions may be used, but there are pitfalls to be aware of:

o Bridges are seldom oriented in a true north-south or east-west direction so


compass directions are of questionable value.

o The railroad often has an established direction of say “railroad east” with the east
always being the lower milepost, whether it’s actually east or not. Using a
railroad direction is manageable, but decisions have been made to change railroad
directions, and it makes it very difficult to compare current reports with historical
data.

o There is a tendency to use identifiers such as “northwest bearing”. This can be


very confusing if the railroad east end of the bridge is not the compass east end. It
would be clearer to say west end, north bearing which would require anyone
reading the report to establish the railroad west end (whether it’s actually compass
west or not) and then find the north bearing.

o Directions can often be used very successfully in conjunction with numbering


systems to get down to smaller details, such as “east bottom flange of #5 floor
beam”.

It is recommended that a written standard identification system be created and provided to


anyone preparing or reading an inspection report. It is also recommended that a key be placed on
the report itself, explaining the numbering system so that each report can stand alone.

Certain inventory information should be provided for each structure. As a minimum, the
following information should be provided:
• The railroad’s name, division and/or subdivision.

• The nature of the crossing (river, roadway, etc.) and name if available.

• A unique identification number as approved by the owner, the age of structure, if it is an


open deck or ballast deck bridge, and the total length, maximum height and number of
panels (spans).

• Name of inspector and members of inspection party, and date of inspection.

Condition Ratings

In addition to the detailed notes describing specific conditions or deficiencies, it is common


practice to assign a condition rating to each condition. Condition rating systems can be very
complex or very simple. An example of a relatively straightforward system would be:

P1 – Requires immediate attention


P2 – Poor condition, keep under observation until repaired
P3 – Fair condition, should be monitored
P4 – Item noted, but of no concern

The rating system should be tailored to the needs of the individual railroad. The rating system
should be clearly defined and provided with each inspection report.

Figure 5-1 Page 1 of 2


Figure 5-1 Page 2 of 2
Figure 5-2 Typical Bridge Abutment
Figure 5-3 Pier Types Sheet
Figure 5-4 Concrete Slab

Figure 5-5 Examples of Arch Bridges


Figure 5-6 Pile Trestle

Figure 5-7 Framed Trestle


Figure 5-8
Figure 5-9 Typical Beam Span Bridge
Figure 5-10 Typical Deck Girder Bridge
(Figure 15-7-3 from AREMA Manual for Railway Engineering)
Figure 5-11 Typical Through Girder Bridge
(Figure 15-7-2 from AREMA Manual for Railway Engineering)
Figure 5-12 Page 1 of 2
Figure 5-12 Page 2 of 2
Figure 5-13 Typical Through Truss Bridge
(Figure 15-7-1 from AREMA Manual for Railway Engineering)
Figure 5-14
ζ

CHAPTER 6

BRIDGE DECKS

General

The deck of a railroad bridge is essentially an extension of the track structure over the bridge.
Like the track structure on a railroad roadbed, the deck of a railroad bridge has several purposes:

• maintain the gauge of the rail


• maintain the line and surface of the rail
• distribute the load of the train to the structure below.

The three main categories of railroad bridge decks are:

• ballast decks
• open decks
• direct fixation decks

Of these three categories, ballast and open deck bridges are the most common, with direct
fixation decks only being used where tight clearances or other unusual factors are present.
The role of the deck as the interface between the train and the superstructure is crucial. Subjected
to the direct action of the train, railroad bridge decks often develop defects faster and require
more maintenance than the rest of the bridge structure. Furthermore, less apparent defects in the
main structure of the bridge often result in more apparent defects in the bridge deck. In this
fashion, a deck defect may indicate a more critical condition in the main structure. For these
reasons it is important that the deck of every railroad bridge be thoroughly inspected on a regular
basis.

This section of the Bridge Inspection Handbook first describes the three main categories of
railroad bridge decks and then reviews what defects to look for while inspecting railroad bridge
decks. An inspection checklist is provided at the end of the section to aid the inspector.

In this sub-section each main category of bridge deck will be defined, the typical components of
each deck category identified, and advantages and disadvantages of each deck category
discussed.

Figure 6-1 Ballast Deck


Ballast Deck

On a ballast deck bridge, the rail is fastened to standard track ties resting in a bed of ballast that
is confined by a ballast pan or tub that is connected to the bridge structure. On some bridges the
ballast tub is an integral part of the bridge structure and is responsible for carrying some of the
shear and bending forces as well as containing the ballast section. This is typically the case in
composite steel bridges and concrete box, tee, and slab bridges. On many steel and timber
bridges, the ballast tub or deck only holds the ballast section and transfers the load to the main
structure below without contributing any structural strength to the span.
Ballast deck bridges have several advantages and are the preferred type of deck on high traffic
and high speed lines. The ballast section has excellent dampening characteristics when
compared to the other deck types. A properly constructed ballast deck reduces the impact loads
on the structure, so the shock from bad rail joints or flat wheels will not do as much harm to the
main structure. The line and surface on a ballast deck bridge are easier to maintain than other
decks and much of the equipment used to maintain the track off the bridge can be used on the
bridge. Adjusting for bridge settlement and small track raises is an easier task with a ballast deck
bridge because reasonable amounts of ballast can be added without any adverse effects.

One of the biggest disadvantages of a ballast deck is the added weight of the ballast on the
structure. More material has to be used for the main structure of the bridge to carry the
additional weight of the ballast. Furthermore, the added weight of ballast and additional material
for the main structure can make constructing and maintaining the bridge more difficult, and more
expensive. Another disadvantage of a ballast deck is that the ballast section can retain moisture
if the deck is not properly drained. This moisture can result in premature deterioration of the
main structure of the bridge. Finally, the depth of ballast deck is typically greater than other
deck types, limiting the use of ballast decks when overhead clearance is tight.

The sketches below show the various components of a ballast deck structure. These drawings
show three basic ballast deck layouts. There are many ways to configure a ballast deck and the
details can vary greatly, but these three examples provide a good introduction to the ballast deck.
Another example is steel deck plate on closely spaced transverse beams.

A
A
F

B E F

B E
D
D
G H

Timber Ballast Deck on Timber Stringers

Timber Ballast Deck on a Steel Span


A Track Structure
A
B Ballast

C Ballast Slab

F B D Ballast Planks

C E Ballast Retainer Straps

F Ballast Retainer
G
G Girders or Stringers

H Timber Stringers

Concrete Slab Ballast Deck on a Steel Span

Figure 6-2 Examples of ballast deck structures


Figure 6-3 Open Deck
Open Deck

An open deck is a bridge deck where the rail and tie plate is fastened to timber bridge ties that
rest directly on the main structure of the bridge. Open decks are typically used on steel or timber
spans. Open deck concrete bridges are uncommon. Unlike some ballast deck structures, an open
deck does not contribute to the strength of the span. Ties on many open decks are subjected to
large bending and shear forces (see section “Open Deck w/Walkway on a Steel Bridge” below).
The chief advantage of an open deck is that it is light and easy to install, making open decks
common on lighter density rail lines and on long span bridges where weight is a concern.
Typically, the installation cost for an open deck is less than both ballast decks and direct fixation
decks. Because of their lower weight, less material is required for the main structural members
of an open deck bridge, further reducing the initial construction cost. Open decks are also
advantageous when overhead clearance issues exist, since the depth of an open deck is typically
less than a ballast deck.
One of the greatest disadvantages of an open deck is that it is difficult to adjust the line and
surface of the track on the bridge. Furthermore, an open deck does not dampen the impact of a
moving train as well as a ballast deck.
The figure below shows the typical components of an open deck on a timber bridge.
Open Deck on a Timber Bridge Open Deck w/ Walkway on a Steel Bridge
Guard Timber

Rail
Outrigger Tie
Tie Plate
Hand Rail Bracket
Bridge Tie
Hook Bolt
Drive Lag

Boat Spike

Stringers Figure 6-4 Open Decks


Figure 6-5 Direct Fixation
Direct Fixation Deck

On a direct fixation deck, the rails are anchored directly to the main structure of the bridge.
Direct fixation decks are not very common. They are typically used when overhead or underside
clearance is tight. They offer a very good line and surface if properly installed and maintained.
However, direct fixation decks may be more difficult to install and maintain than more
conventional decks. A disadvantage is the inability to mitigate large impact forces.

It is difficult to provide an extensive review of direct fixation deck inspection because most
systems are proprietary and the designs vary greatly. The information in this handbook covers a
few basic points of direct fixation decks. If you are responsible for a bridge with a direct fixation
deck, the best approach is to contact the manufacturer of the system for the proper inspection and
maintenance procedures.

The figure below shows one type of direct fixation deck.

Rail

Clip Anchor (uninsulated)

Fastener

Shim Plate

Main Girder

Figure 6-6 One type of direct fixation deck


Deck Defects

There are several ways in which a bridge deck can develop defects. It is difficult to present
every possible defect within the space available in this handbook. The following is a listing of
the most common deck defects which provides an introduction, but is by no means all inclusive.

Line, Surface, and Gage

No matter what type of deck is being inspected, the quality of the track line, surface, and gage
should be the primary criteria to evaluate the condition of the deck. As stated above, the primary
purpose of a bridge deck is to maintain the line, surface, and gage of the track over the bridge.
Loss of any of these three attributes should raise concerns and may indicate the deck or the main
structure has developed a defect. It is important to note that the loss of gage, line, or surface may
be the result of a defect in the deck itself or of a condition in the main structure of the bridge.

Every time a bridge is inspected, the line, surface and gage of the track, should be checked.
When checking the line of the deck, any track swings or kinks on tangent track should be noted.
On curved track, any changes from the design curvature or superelevation should be noted.
Inspection of the surface should note high and low spots in the deck, as well as any areas of poor
cross level. Poor line and surface is often a good indicator of a critical condition with the main
structure, such as settlement. The location of line and surface defects should be noted prior to
inspecting the remainder of the structure.

Rail

Inspection of the rail on a bridge shall be done by a qualified track inspector in accordance with
FRA or Transport Canada regulations. FRA and AREMA have extensive reference material
dealing with rail inspection. Therefore, rail inspection will not be covered in depth in this
section. In some situations a defect in the deck or main structure will result in unusual wear on
the rail. If unusual rail wear is found, its location should be noted prior to inspecting the
remainder of the structure so the inspector can search for a possible root cause.

Open Deck Bridges

Decaying Ties

Decay occurs when the cellular structure of the timber tie is


consumed by a fungus, reducing the strength of the tie. Decay
usually attacks the untreated heartwood of the tie first, often
around spike holes or checks in the treated exterior of the tie. As
decay sets in, the ties will show signs of plate cutting, an outward
sign of an internal decay problem. The ties begin to crush under
the rail, the track spikes will loosen and the track surface and
gauge may be affected. If the ties on a bridge are consistently
plate cut across an area of the span, it is likely that decay has set Figure 6-7 Decaying Ties
in around the spike holes and rail bearing area.
Fire Damage

Located at the top of a structure and evenly spaced, open deck


ties provide an excellent source of well-drafted fuel to be
consumed once a fire is started. Hot spots resulting from
superheated rail and structural steel often cause severe damage in Need Burnt Bridge Photo
the critical bearing area of the ties. Mechanical damage to the
deck also can occur when the superheated rail buckles, pulling
out spikes and pulling the deck away from the bridge.

Mechanical Damage Figure 6-8 Fire Damage

Mechanical damage occurs when ties are subjected to repeated


overloading or large impact loads. Mechanical damage can take
several forms, such as plate cutting, broken ties, and derailment
damage.

Unlike plate cutting due to decay, mechanical plate cutting is


often isolated to a small area of the deck, typically around the
ends of the bridge or joints. Mechanical plate cutting is caused
by excessive pressure on the tie and is typically due to external
Figure 6-9 Mechanical Damage
defects such as low approaches or loose joints.

Broken ties can be the result of dragging equipment, running


joints (see Figure 6-10), excessive loads, or undersize ties. As
with mechanical plate cutting, broken ties may be an indication
of defects beyond the tie itself, such as poor rail anchorage, loose
joints, or excessive loading.

Derailment damage can vary greatly, from a series of small


creases on the surface of the ties to complete destruction of the
deck. After any derailment, a bridge deck should be checked for
broken ties, damaged spikes and plates, skewed ties, damage Figure 6-10 Broken Ties
guard timbers, and loss of gauge.

Skewing & Sliding Ties

The ties on an open deck should be evenly spaced and square to


the rail. Sliding ties will move apart or along the bridge, and
skewed ties will become out of square to the rail. The ties of an
open deck will slide and skew if they are not properly attached to
the main structure of the bridge or secured by an effective tie
spacer.

Sliding ties pose a concern to both trains and workers walking on


the bridge deck. Sliding ties can open large gaps in the deck, Figure 6-11 Skewing
& Sliding Ties
leaving the rail without proper support and opening holes that a
worker walking on the deck may fall through. Skewing ties will
affect track gauge and lead to mechanical damage to the ties, tie
plates and the base of the rail.

Checked and Split Ties

A check is the separation of wood fibers caused by shrinkage as


the wood dries. Checks are typically not a problem, but do provide
an avenue for moisture and debris to penetrate and further degrade
the wood. A split is the complete separation of wood extending
from one face of the tie to another. Ties that were not dried
sufficiently prior to treatment or ties with poor grain can severely
split when overloaded or subjected to changes in temperature and
humidity. Splitting ties can result in loosening track spikes, loss of
lateral and longitudinal restraint of the track, and offer a place for
Figure 6-12 Checked &
decay to set in.
Split Ties
Whether a split tie is still effective is dependant on the size and
location of the split. Any tie with a split that passes under the tie
plate area and out the end or side of the tie should be considered
ineffective.

SPECIAL NOTE

A A
B B

Open Deck Ties on Steel Structures

There is an additional concern with open decks on steel bridges with only two
stringers or beams running under the deck. For reasons of stability the two stringers
or beams in many steel bridges are set wider than the gauge of the rail. This means
that there are no structural members directly under the rail (See “A” rail above). With
no support directly under the rail, the tie must carry the load of the train in bending
and shear out to the main structural member. There is no redundancy in this system,
and so development of defects in a series of ties can result in a sudden loss of surface
and elevation (See “B” above).
Ballast Deck Bridges

Ballast Leak

Decay, mechanical damage and corrosion can create holes in


ballast decks resulting in ballast loss. Ballast leaks undermine
the ballast support of the track structure, resulting in low spots in
the track and poor cross level. Ballast leaks are usually not
immediately apparent from the top of the bridge making it
important to inspect the underside of a ballast deck to spot holes.

Ballast Retainer Defects Figure 6-13 Ballast Leak

Ballast retainers are critical in maintaining the ballast section on


a ballast deck bridge. Loss of the ballast retainer may lead to the
undermining of the ballast shoulders and a decrease in track
stability.

The following is a review of common ballast retainer defects:

ƒ Decayed Ballast Retainer – Timber ballast retainers can


decay and eventually fail. As internal decay progresses
the bolts securing the retainer will loosen and the retainer
will begin to deflect and crush under the weight of the Figure 6-14 Decayed Retainer
ballast. As with many ballast deck defects, decay of
timber ballast retainers is accelerated by moisture trapped
due to poor drainage.

ƒ Overloaded Ballast Retainer – Timber and concrete


ballast retainers often fail when subjected to excessive
ballast load. As more ballast is added to the span the
ballast retainer becomes overloaded, causing it to roll out
and away from the track, eventually leading to failure.

ƒ Cracked & Heavily Spalled Retainers – Poor concrete or


concrete that has been subjected to excessive moisture
will spall and develop cracks over time. Mild pattern
cracking and spalling typically does not present a threat
to the safe operation of trains. However, spalling around
deck joints and on ballast curbs will eventually progress
to the point that ballast loss will occur, and should be
monitored or repaired.
Fire Damage

Timber ballast decks can be adversely affected by a fire. Typically, a fire will only char the
underside of the timbers because the deck timbers are tightly packed and buried in ballast,
starving the fire of oxygen. Hot spots can develop, however, and the underside of the deck
should be checked for ballast leaks and excessive charring that may reduce the bending capacity
of the deck timber.

Beyond the deck timbers, fire can also damage ballast retainers, typically even more so than the
deck timbers. The two most common ways fire can damage a ballast timber are excessive
section loss due to fire char and the weakening of the fasteners securing the retainer to the deck.
In the former case, the fire burns through the retainer allowing ballast to leak out of the sides of
the deck. In the latter case, the fire superheats the steel fasteners holding the retainer in place,
softening the steel and charring out the holes the fastener passes through causing the retainer to
roll out.

Excessive Ballast Load

One of the advantages of a ballast deck is that reasonable amounts of ballast can be added to
correct track surface defects. Unfortunately, over time too much ballast can be added to the deck,
increasing the load on the main structure below. Typically, ballast deck bridges are designed
with 8” to 18” of ballast under the ties. Excessive ballast beyond the design ballast depth
increases the stresses in the ballast pan, the main structure, and the substructure of a bridge.

The additional ballast on a structure may be distributed evenly across a bridge as a result of a
track raise, or in some cases additional ballast is heaped at a low spot created by settlement of an
area of the bridge. In the former case, the inspector should note the depth of ballast and have the
structure checked to see if additional stresses are significant. In the latter case, the need for the
excess ballast is the result of a defect elsewhere in the bridge and it is important to determine
what defect is causing the low spot.

Fouled Ballast

Over time, ballast rock degrades and foreign material accumulates in a ballast pan fouling the
ballast with fines, silt and mud. Fouled ballast traps moisture, accelerating decay, deterioration,
and corrosion of the ballast pan, track structure, and the main structure of the bridge.

Poor Drainage

Poor drainage is a major contributing factor in many ballast deck defects, including fouled
ballast, tie and deck decay, spalling concrete, and corrosion. There are several reasons why a
deck does not drain properly. The deck may have been designed with insufficient drainage, or
the original waterproofing and drainage may have been improperly installed. Even properly
designed and installed drainage will become clogged and cease to work effectively over time,
usually before the rest of the bridge has reached the end of its useful life.
Unfortunately, there is no good way to inspect the waterproofing and drains since the drainage
system is usually buried under ballast. The first signs of poor drainage are typically defects
caused by poor drainage such as efflorescence on the underside of the concrete deck or corrosion
on a steel ballast pan. An inspector can check deck drains immediately following a rain storm for
water, but even this is imperfect because water can still be pooling elsewhere on the deck.

It is important to determine if a defect is caused by poor drainage during an inspection of a


ballast deck. If resources are to be spent repairing a defect without correcting the poor drainage
causing the defect, the repair may not be effective.

Direct Fixation Deck Bridges

Direct fixation fasteners should be inspected carefully for impact damage. Such damage can
include cracking of fastener components, including welds and bolts. Damage can also occur to
the superstructure itself in the area of the fasteners, particularly if they are welded to the
superstructure.

Another item to observe on direct fixation is the condition of the insulators where required for
track circuits. Wear in these insulators can cause a shunt in the track circuit, causing false signal
indications. When signal circuit failures due to poor insulators start to occur, it is often time to
replace all insulators unless a specific condition is causing the failure in an isolated location.

Inner Guard Rails

Where inner guard rails are provided, spacing between the guard rail and running rail should be
uniform. Guard rails should be adequately fastened to the ties or deck. All sections of guard rail
must be properly bolted together. The top of the guard rail must not be higher than the top of the
running rail. Guard rails should be terminated on the bridge approach with a nose or point
section.

Deck Inspection Checklist

Line
__ Are there any alignment deviations in the track? If so, where and to what degree?
__ Is the track properly centered on the bridge?
__ Is the track on the approach lined well to the track on the bridge?
__ Are the items noted above due to a deck defect or a defect in the main structure?

Surface
__ Are there any low spots? High spots? If so, where and to what degree?
__ Are the any areas out of cross-level? If so, where and to what degree?
__ Are the approaches low? High? If so, how much?
__Are the items noted above due to a deck defect or a defect in the main
structure?

Guard Rail
__ Is it present or required?
__ Is it securely fastened to the deck and properly bolted?
__ Is it the proper size?
__ Is the spacing uniform?
__ Is the rail section the proper weight?

Gauge
__ Are there any areas of tight or wide gauge? If so, where and to what degree?

Rail
__ Are there any unusual areas of wear? If so, where and what is the apparent cause?
__ Are the joints tight and fully bolted?
__ Is the rail well anchored on the bridge and on the approaches?

Open Decks
__ Tie Condition
Split ties
Broken ties
Decayed ties
Derailment or fire damage
Slid or skewed ties
Horizontal shear cracks

__ Fasteners
Loose or missing track spikes
Loose or missing rail anchors
Loose or missing hook bolts (on steel spans)
Insufficient or non-existent deck to stringer fasteners (on timber spans)
Loose or missing tie spacer (guard timber)

Ballast Decks
__ Tie Condition
Decayed ties
Derailment damage

__ Ballast Condition
Fouled ballast
Excessive ballast
Ballast leaks
Evidence of poor drainage

__ Ballast Pan or floor


Holes in the pan
Fire damage
Spalling, corrosion, or decay
Loose or missing fasteners
Evidence of poor drainage or loss of waterproofing
__ Ballast Retainers
Overloaded ballast retainers
Decayed or damaged ballast retainers
Fire damage
Spalling or cracked concrete
Mechanical damage

__ Fasteners
Loose or missing track spikes
Loose or missing rail anchors
Loose or missing hook bolts securing deck panels (on steel spans)
Loose or missing bolts for ballast retainers

Direct Fixation Decks


__ Fasteners
Loose or missing rail anchors
Loose or missing anchor points

__ Insulation
Loose, damage or missing rail insulators

__ Substrate
Spalled or deteriorated grout/concrete
Pack rust, cracking, or delamination.

Walkways
__ Loose or missing walkway planks
__ High planks or other tripping hazards
__ Debris on the walkway
__ Security of handrail posts
__ Handrail or wire rope condition
__ Anchorage and tension of wire rope

Emergency
If the bridge inspector encounters a bridge condition that affects the integrity
of the bridge under train loads, contact the railroad dispatcher and
responsible authority to stop any trains and arrange for immediate repairs.
Criteria for the limits of acceptable/tolerable conditions must be provided by
the Railroad Owner or their designated engineer based on the principles of
acceptable rating contained in the appropriate chapter of the AREMA
Manual for Railway Engineering.
CHAPTER 7

TIMBER BRIDGES

General

In the early years of railroad construction, timber was used extensively due to availability and the
relative ease of construction. Although, for a variety of reasons, timber is not often the bridge
material of choice anymore, there are thousands of timber railroad bridges in use throughout the
country that will require inspection and maintenance for many years to come.

Timber bridges should be inspected at least once a year. Bridges with any noted progressive or
serious deterioration should be inspected on a more frequent basis and the results of these special
inspections should be maintained in the same manner as the routine inspections.

Material Properties

The material properties for timber vary with the species being used and that selection is, at least to
some degree, based on the species native to any region of the country. Just a few examples
include hardwoods in the east and northeast, pine in the south and southeast and fir in the west.
Many of the specific properties considered in the design of timber bridges are related to the
density of the particular species and grade, which is determined by the basic cell structure of the
wood. Factors affected by the grade of lumber are strength, durability, resistance to decay, and
the ability to accept preservative treatment.

Strength

The strength of any piece of lumber is determined by the nature of the applied loads and
orientation of the member. Factors that affect the strength include how the wood was sawn, knots,
direction and slope of grain, moisture content, distortion, and splits, checks and shakes. These are
all factors that are considered in the grading of structural lumber. The unique properties of wood
provide valuable insight on likely areas of deterioration to be evaluated during inspections.
ƒ Compression – The ability of wood to resist compressive forces differs significantly
depending on whether it is loaded parallel or perpendicular to the grain structure. Wood
loaded parallel to the grain, such as a pile, has a much greater capacity per unit area than
a cap, which is loaded perpendicular to the grain.
ƒ Tension – Wood is not typically used in pure tension applications, but exhibits good
strength in tension when bending moment is applied, such as in stringers.
ƒ Bending – Wood is quite strong in bending provided there are no major defects, such as
large knots, that break up the normal grain structure.
ƒ Horizontal Shear – Horizontal shear, which is the tearing of the wood material parallel to
the grain, is found in members such as stringers which are subject to large bending forces
(open deck ties may also experience considerable horizontal shear forces depending on
the support configuration).
Durability

Wood can be quite durable, but that durability depends on the quality of the wood and
preservative treatment, the natural environment, and the applied loadings. Wood with a low
moisture content and good preservative treatment is more resistant to insect and decay damage.
Bridges in an arid environment such as the southwest are much less susceptible to decay than
those in a warm, moist environment such as the southeast. And timber structures hold up very
well mechanically under the loads they were designed to carry.

Wood can also perform surprisingly well in fires. A coat of charred material will often form on
the outside that protects the remainder of the wood from further fire damage.

Bridge Notes

Timber bridges have a large number of individual components that interact with each other.
There are often a significant number of deficiencies found due to normal mechanical wear and
decay. The bridge inspector’s notes for each bridge are, by necessity, very detailed and shall be
written while at the structure after a careful examination has been made. Inspection records for
each individual bridge structure must include the following information (in addition to the
inspection findings):

ƒ The railroad’s name, division and/or subdivision.


ƒ A unique identification number as approved by the owner, the age of structure, if it is an
open deck or ballast deck bridge, and the total length, maximum height and number of
panels (spans).
ƒ Name of inspector and members of inspection party, and date of inspection.
ƒ Number of bents, towers, spans or panels in each bridge in the direction in which the mile
post numbers increase, starting with the dump bent as No. 1. Number the piles in each
bent or tower and the stringers in each panel from left to right, when facing in the
direction in which the mile post numbers increase.

Modes of Timber Deterioration

Stress/Mechanical Deterioration:

Stress/Mechanical wear is caused by train loads. Examples of this type of wear can be found at
points of bearing such as stringers to caps and caps to piles. Shims or gaps found at the tops of
piles or under stringers are good indications of mechanical wear problems. Crushing of the
timber at bearing points, or individual pile settlement is evidence of stress/mechanical wear.
These conditions are progressive and will cause movement throughout the structure at the
passing of every train. See Figure 7-1.

Other types of mechanical wear include abrasion from ice flows or sediment transport. Impact
damage can be caused by highway loads, floating or submerged debris, watercraft, derailment,
etc.
Organic Decay:

Organic timber decay can be found most frequently around connections using bolts, drift pins,
and deck fasteners. These areas can be found at:
ƒ cross braces to piles (See Figure 7-2.)
ƒ cross braces to caps
ƒ caps to piles
ƒ stringers to caps (See Figure 7-3.)
ƒ deck connections to stringers (See Figure 7-4.)
ƒ tie plates
ƒ ballast timber connections to deck boards and outside stringers

Organic decay usually starts internally and cannot be found unless an inspector “sounds” the
wood at these areas. When a dull “watermelon” type sound is heard, there is a good chance that
the timber has a decay void. In addition to points of connection, external organic decay is
predominate around:
ƒ piles at the ground line (See Figure 7-5.)
ƒ tops of caps at stringer connections
ƒ across the tops of stringers between the ties or ballast deck planks

All of these locations are subject to frequent wet-to-dry conditions. If a void is not discovered
through inspection, and the outer shell thickness grows too thin, it is possible to experience a
sudden failure of a timber member under train traffic. See Figures 7-6 & 7-7.

Insects:

Timber bridge components are subject to attack from insects such as:
ƒ termites
ƒ carpenter ants
ƒ powder-post beetles
ƒ various types of borers

Although these types of insects are not always easily found, there are indications of their
presence. White mud shelter tubes or runways extending up from the earth to the wood and on
the sides of masonry substructures are signs of infestation of termites. The outer surface of
timber piles with surface pocks and small holes may be signs of beetles. An accumulation of
sawdust on the ground is a good indicator of carpenter ants. If the timber bridge is located in
water, especially around ports near the ocean, the inspector should become familiar with the
indications of shipworms, marine borers, and crustacean borers. A competent diver may be
required to identify these types of infestations.

Collision:

Timber bridges located over roadways are subject to damage from high or wide loads. Due to
the relative short span lengths of timber bridges, the bents are subject to collision damage from
debris floating in the waterways. The inspector needs to be alert to this type of damage and note
such damage in the inspection records. See Figures 7-8 & 7-9.

Inspection Methodology

The method of inspecting timber, regardless of its location in the structure should be as follows:

ƒ Make a careful surface inspection of physical characteristics of each timber component.


These observations need to be focused on surface bulges, cross grain splits, and tension,
horizontal, or shear failures that may have developed from uneven bearing, original
defects, overstress, organic decay, or other causes. Note whether the timber bridge
components are treated or untreated.
ƒ Test each timber and pile for soundness, especially at points of contact with other
timbers, ground, at the high and low water line, and where end grain bears on a sill or
cap.

Tests should be made by sounding with the knob end of an inspection bar or lightweight
hammer, using care to avoid injuring or disfiguring the fiber. If hollow or dead sound results,
determine nature and extent of the defect by drilling, or boring with an increment borer.
Measure and note void and shell dimensions as well as identifying the pile and location of the
test hole. When the measurements and inspections are complete, carefully plug the drilled hole
with a treated wood dowel.

Make a careful surface inspection of the timber and adjacent ground surface for evidence of
termites, carpenter ants, marine borers or other destructive insects.

Observe the timber condition at and above the water surface. The water elevation often varies
and this type of wet-to-dry cycle will create or accelerate timber decay. If the timber bents
cannot be reached during routine inspections due to constant water conditions, such as in a pond,
lake, or wide stream, special inspections should be scheduled using a boat to inspect the timber
bents.

Make inspection of new work. Where timber is field cut, check for exposed untreated wood.
Inspect to see that all cuts are square and that members fit together snugly, with full bearing to
transfer loads. Check fasteners for tightness.

Details of Inspection

The bridge inspector’s notes for each bridge shall be written while at the structure after a careful
examination has been made covering the following points:

Environment

ƒ Waterway: Measure and record high water mark if obtainable. Timber bridges often
replace a previous generation structure. Look for old pile stubs that might impede the
flow of water and contribute to scour problems. Accumulated drift is particularly
troublesome on timber structures causing such hazards as potential fire, scour and
erosion, and backing up water and damaging approach rail embankments or adjacent
property. Severe drift accumulation can also cause the bridge and track to be pushed out
of alignment during high flows. Note if there is an excessive distance from a bent’s
bottom sash brace, or cross brace, to the ground line as an indication of scour that will
cause loss of pile penetration. See Figures 7-10, 7-11, 7-12, 7-13, 7-14, 7-15, 7-16, 7-17
& 7-22.

ƒ Roadway: Where highways or roads pass under the structure, note their location by bent
numbers, state class and name of road and skew of the roadway. Should there be
evidence that vehicles have impacted the structure so as to seriously damage or weaken
it, prompt report should be made.

Fire

Note if the bridge has evidence of any previous fire damage, and if so, where the damage is
located and the extent and depth of char. Other inspection observations to be made should
include:

ƒ Note whether surface of the ground around and beneath the structure is kept clean of
grass, weeds, drift or other combustible material.
ƒ Where sheet metal is used as a fire protection covering for deck members, note condition
of metal and fastenings. See Figures 7-18, 7-19, 7-20, 7-21 & 7-23.
ƒ Note if any other method of fire protection has been used, such as fire retardant salts,
external or surface protective coatings, fire walls, or fire wires to alarms. Record such
apparent observations as are pertinent to the physical condition and effectiveness of such
protective applications.
ƒ Where water barrels are provided, note the number, condition, if filled, and if buckets for
bailing are on hand. If sand is used, note whether bins are full and in condition to keep
the sand dry.
ƒ Note if timber, particularly top surfaces of ties and stringers in open deck bridges, is free
from frayed fiber, punk wood, or numerous checks that would make the structure more
susceptible to fire.

Track

Closely observe the track from both ends of the structure, and, if possible from a distance on
both sides of the structure. Observations should be made of the following:

ƒ State whether track is level or on a grade, and if alignment is tangent or curved. If on a


curve, note how super elevation is provided, whether by cutoff in the bents, taper in the
caps, shims under the tie plates, or in the ballast section. Note location of track centerline
with reference to the centerline of chords. See Figure 7-24.
ƒ Observe condition of track embankment at the bridge ends for fullness of crown,
steepness of slopes and depth of bulkheads. Note whether track ties are fully ballasted
and well-bedded.
ƒ Record the weight and condition of the running rails and inside guard rails, if in place;
also the condition of the rail joints and fastenings. Note the size and condition of the tie
plates. Also note whether rail anchors are installed on the bridge and/or approaches.
ƒ Where track is out of line or surface, measure and record the location, amount and
probable cause.

Deck

Observe the deck from the top, bottom, and sides. Observations should be made of the following:

ƒ If the bridge has a walkway and/or a handrail, note the method of construction, condition,
and the location, if on the right, left, or both sides of the bridge.
ƒ If an open deck bridge, note the width, height, and length of the ties and the tie count and
spacing. Note the size, type of material, and method of anchorage used for control of tie
spacing. Note condition as to soundness, mechanical wear, spike killing, plate cut, and
other defects. See Figures 7-25, 7-52 & 7-53.
ƒ If ballast deck bridge, note the size, type, and condition of the ballast deck and retainers.
Note the height, type, and condition of additional ballast retainers that have been added
due to track raises. Note if ballast is clean and in full section, or if the ballast is fouled
with mud or fines. See Figures 7-26, 7-50 & 7-51.
ƒ Inspect condition of any motorcar setouts or man refugees. See Figure 7-27.

Superstructure

Closely examine the superstructure and observe the following:

ƒ Examine all stringers for soundness and surface defects. Note size and kind, and the
number used in each panel. Note if bearing is sound and uniform, if all stringers are
properly chorded and securely anchored, and if all shims and blocking are properly
installed. Note whether packers or separators are used and the condition of all chord
bolts. Note if a multiple span bridge is constructed using continuous stringers or simple-
span stringers.
ƒ Note any horizontal shear cracks which are cracks that develop at mid-depth on the
stringers and extend the entire length of one or more spans. Inspect large knots closely
for any evidence of cracks or splits.
ƒ Look closely at any crushing of the stringers over caps. Identify if the crushing is due to
mechanical wear or due to decay on the inside of the stringer, causing it to crush.
ƒ Note if any notches have been placed at the ends of the stringers to provide uniform
bearing over the caps. If there are notches, note if they are cut square or otherwise.
Carefully observe the ends of stringers, especially at any notches, for indications of
longitudinal splits. See Figures 7-28, 7-29, 7-30, 7-32 & 7-33.
ƒ Note method of anchoring stringers to the bents and the condition of the anchorage
system and note indications of movement at the bearing areas. Check carefully to see that
the parting lines where stringers butt up against each other are centered over the caps.
See Figures 7-31 & 7-54.
Substructure

For the substructure, closely observe the following:

ƒ Measure and record the distance from base of rail to ground line at each bent.
ƒ Record number and kind of piles or posts in the bents or towers. Note uniformity of
spacing and the location of any stubbed or spliced members, especially if the bridge is on
a curve or the bent is tall. Note any piles that have been “posted” and type of connection.
ƒ Make examination of all piles and posts for soundness, noting particularly the condition
at points of contact with the caps, girts, bracing, sills, and at the ground or water line.
Make note of the number and depth of shims, or open gaps between points of contact.
Where timber caps have been replaced with concrete caps, closely observe the outside
batter piles, since they may now be experiencing higher loads due to the stiffer cap. See
Figures 7-34, 7-35, 7-36, 7-37 & 7-38.
ƒ Examine all bents and towers for plumbness, pumping, settlement, sliding and racking,
and give an accurate description of the nature and extent of any irregularities. Note
particularly whether top-caps, intermediate-caps, and sills have full and uniform bearing
on the supports. See Figures 7-44, 7-45, 7-46, 7-47, 7-48 & 7-49.
ƒ Ascertain whether bents and towers are sway, sash and tower braced, and if girts and
struts are installed.
ƒ Indicate the condition and method of support for framed bents, such as concrete
pedestals, cut-off piles, and concrete footings.
ƒ For timber caps, inspect for splits, ring separations, bulging and crushing. For concrete
caps, inspect for splits in the concrete and for settlement or signs of failure at piles. If a
concrete cap shows signs of breakage, look at the ends of the cap for any indication that
the reinforcing strands are slipping and not properly bonded to the concrete. See Figures
7-39, 7-40, 7-41, 7-42 & 7-43.
ƒ Examine all fastening devices for physical condition and tightness.
ƒ Observe action of bridge under movement of trains, where practicable, in order to
evaluate better the riding condition and stability of the structure.

Effects of Unit Trains on Open Deck Timber Bridges

On railroads that operate unit trains (trains with a large consist of identical loaded cars), the
longitudinal forces exerted on open deck timber bridges often cause damage which may manifest
itself in:

ƒ Excessive longitudinal movement of the rail in the tie plates, usually in the direction of
the predominant loaded unit train movement. On smaller sizes of bolted rail, where angle
bars overhang the base of the rail, there is often considerable tie damage when the angle
bar catches on adjacent track spikes, splitting the ties.
ƒ Longitudinal movement of the stringers (and possibly the rest of the bridge structure) in
the opposite direction of the predominant loaded unit train movement, causing one or
more of the following conditions, depending on the relative stiffness and quality of
connections on individual timber members and bents:
o Stringer parting lines moving off the center of the cap
o Caps splitting through the drift pins
o Caps wanting to roll off of piles
o Tops of piles splitting at drift pins
o Bents leaning longitudinally along the track centerline (more common on taller
bents and particularly frame bents)
o A wallowing out of the ground line at the base of the bent piles
o Stringers pulling off of the end bent cap on the trailing end of the unit train
movement, and pushing into or through the backwall timbers on the leading end.

Finding one or more of the above conditions at an isolated bent does not indicate a serious
longitudinal force issue, but if these conditions exist at a number of bents, compile a listing of
these defects by bent so that a more thorough evaluation can be made regarding bridge stability.

Note: If it appears that there is significant longitudinal movement, try to observe the bridge under
unit train load. Relative to a fixed point, if there is an issue, the rail will typically move in the
same direction as the train while the bridge structure simultaneously moves in the opposite
direction, often times with several inches of relative movement. Much of the movement will
spring back after the passage of the train, but the movement and resultant damage tends to be
cumulative.

TIMBER BRIDGE INSPECTION CHECKLIST

Track

ƒ Track alignment and surface profile. Determine if any defects are track or structure
related.
ƒ Condition of rail, fasteners, and other track material (OTM)
ƒ Condition of track or bridge ties
ƒ Ballast
o Clean or fouled
o Depth
o Adequate shoulder
o Ballast retainers

Walkways

ƒ Walking surface
ƒ Posts and handrails
ƒ Mounting brackets
ƒ Hardware/fasteners

Superstructure (stringers)

ƒ Overall alignment – Vertical and horizontal


ƒ Mechanical damage
ƒ Decay or insect damage
ƒ Crushing at caps
ƒ Horizontal shear cracks
ƒ Cracking or broken stringers

Substructure and Foundation

ƒ Crushing or broken caps


ƒ Connections & shims
ƒ Decay (points of contact & groundline)
ƒ Check bents for:
o Plumb
o Settlement
o Racking
o Pumping
ƒ Scour or erosion
ƒ Bracing
ƒ Note framed bents and posted piles

Emergency

If the inspector discovers a bridge condition that affects the integrity of the
bridge under train loads, contact the railroad dispatcher and responsible
authority to stop any trains and arrange for immediate repairs. Criteria for
the limits of acceptable/tolerable conditions must be provided by the Railroad
Owner or their designated engineer based on the principles of acceptable
rating contained in the appropriate chapter of the AREMA Manual for
Railway Engineering.
Timber Bridges

Figure 7-1 Crushed Cap Figure 7-2 Pile Decay at Brace Bolt

Figure 7-3 Pile Cap Decay Figure 7-4 Stringer Decay at Cap

Figure 7-5 Groundline Decay Figure 7-6 Failed Pile


Figure 7-7 Failed Pile 2 Figure 7-8 Stringer Damaged by Impact

Figure 7-9 Pile Broken by Impact Figure 7-10 Drift

Figure 7-11 Drift 3 Figure 7-12 Vegetation 2


Figure 7-13 Silting Figure 7-14 Pile Cutoffs

Figure 7-15 Scour Figure 7-16 Bank Erosion Loss of Pile Penetration

Figure 7-17 Scour Under Headwall Figure 7-18 Bridge Fire


Figure 7-19 Fire Damage to Deck Figure 7-20 Fire Damage to Piles and Bracing

Figure 7-21 Fire Damage to Tie Deck Figure 7-22 Severe Erosion

Figure 7-23 Tin Covering of Timber Pier Members Figure 7-24 Poor Line & Surface
Figure 7-25 Plate Cut Ties 2 Figure 7-26 Fouled Ballast

Figure 7-27 Lookout Details #3 Figure 7-28 Broken Stringer

Figure 7-29 Broken Stringers Figure 7-30 Stringer Decay at Cap 2


Figure 7-31 Stringer Parting Lines Figure 7-32 Horizontal Shear Crack & Crushing

Figure 7-33 Badly Decayed Stringer Figure 7-34 Badly Failed Piles

Figure 7-35 Chord Shims 2 Figure 7-36 Missing Pile


Figure 7-37 No Pile Penetration Figure 7-38 Pile Shims

Figure 7-39 Broken Cap 2 Figure 7-40 Cap Crushing Under Stringer

Figure 7-41 Failed Cap with Ring Separations Figure 7-42 Cornering – Crushing Cap
Figure 7-43 Cornering – Under Stringer Chord Figure 7-44 Leaning Bent #2

Figure 7-45 Hogjaw Braces Figure 7-46 Settling Bent

Figure 7-47 Settling Bents Figure 7-48 Twisted Bent


Figure 7-49 Crushed Sill Figure 7-50 Excessive Ballast Depth

Figure 7-51 Failed Ballast Timbers Figure 7-52 Damaged Ties at Running Rail Joint

Figure 7-53 Decay Under Tie Plate Figure 7-54 Failing Chord Shims
CHAPTER 8
CONCRETE & MASONRY BRIDGES AND FOUNDATIONS

SECTION 1
CONCRETE PROPERTIES

General

There are many factors that contribute to the deterioration of concrete structures, many of which,
such as original composition and placement, are unknown at the time of inspection. Many
instances of concrete deterioration observed today can be traced back to the specific components,
mixing, and placing methods used at the time of construction. Although concrete has been used
as a building material for a long time, the importance of proper mix proportions, clean
aggregates and water, curing techniques, etc. continue to evolve. As a result, it is important to
have a working knowledge of the basics of concrete construction.

Material Mechanics

Concrete consists primarily of a mixture of cement (almost always Portland Cement), fine and
coarse aggregates, and water. The water and cement chemically react in a process called
hydration, which bonds all the materials together. The resulting mass consists predominantly of
fine and coarse aggregates. The strength and durability of concrete is a function of the mix
proportions, as well as factors such as where it is placed, temperature during placement, and
finishing and curing techniques.

Concrete has five basic characteristics important to bridge construction:

• Strength – Concrete is very strong in compression, but weak in tension, shear, and
torsion.
• Porosity – Concrete is porous, the cement paste does not fully fill the voids around the
aggregate. Consequently, concrete will absorb water.
• Elastic – Under ordinary loads, concrete will deform but is quite brittle.
• Durable – Properly proportioned, mixed, and placed concrete is very durable. Climate
and exposure greatly affect its durability. Durability and strength are decreased when
workability is increased by adding water during placement.
• Fire resistant – Although high quality concrete is relatively resistant to the effects of fire,
intense heat can cause severe damage including cracking and spalling.

As previously mentioned, the compressive strength of concrete is relatively high. Since it is a


somewhat brittle material and the tensile and shear strengths of concrete are relatively low,
reinforcing steel is added to resist tensile and shear forces while the concrete itself carries most
of the compressive forces. Reinforcing steel may also be placed in such a manner as to aid in the
resistance of compressive forces.
Aggregates are classified as fine or coarse and normally occupy 70 – 75% of the volume of
hardened concrete. Fine aggregate is generally sand particles that will pass a No. 4 sieve and
coarse aggregate consists of stone that will be retained on a No. 4 or larger sieve. Chapter 8 of
the AREMA Manual for Railway Engineering establishes recommendations on the maximum
size of coarse aggregate as well as other requirements for reinforced concrete.

Engineering Characteristics

Concrete in Compression

Stress is defined as a resistance to external forces and is measured in terms of force per unit area,
normally pounds per square inch (psi). Strain is a term that refers to the deformation
(elongation or shortening) that occurs in a material as it is loaded to a certain stress.

There is an important relationship between stress and strain in concrete. Stress is proportional to
the external force being applied to a member. As stress increases, strain increases. However,
when the strain (deformation) exceeds a value of about .0025 in/in, the concrete begins to crush
and it is unable to carry additional stress. In fact, the stress begins to decrease with increasing
strain beyond this point. The important thing to note is that once concrete starts to fail in
compression (when there is large deformation or deflection), the ultimate failure can be quite
sudden and catastrophic (brittle failure).

ASTM has standardized testing to determine the compressive strength of concrete. The test
involves compression loading a specimen to the point of failure. Currently, 28 day concrete
compressive strengths range from about 2,500 psi to 9,000 psi with 3,000 to 4,000 being the
most common range for reinforced concrete, and 5,000 to 6,000 psi the most common for
prestressed concrete members. Concrete strength varies with time, and the specified strength is
that which (by design) exists 28 days after placement. Generally, concrete attains approximately
70% of it 28 day strength after 7 days and 85 – 90% in 14 days.

Concrete in Tension

Unreinforced concrete is very weak in tension. In order to give concrete the ability to carry
tensile stresses, reinforcing in the form of steel bars or welded wire fabric is embedded into
conventionally reinforced concrete. While prestressed and post-tensioned concrete also utilize
steel reinforcing bars, the primary reinforcement used to provide bending capacity consists of
steel wire and strands respectively.

Behavior of Reinforced Concrete

Flexural Stresses

Flexural stresses are those which are caused by loads which bend concrete such as in a beam or
slab. Concrete may crack in flexure even under normal loads. These cracks are very small,
hairline cracks and are barely noticeable. Their presence does not indicate any structural
problems in conventionally reinforced concrete. Flexural cracks in prestressed or postensioned
concrete are very significant and are discussed in further detail later in this section. It is
important to record any growth or changes in crack lengths or widths.

If a concrete structure is overloaded, failure is governed by one of two modes, depending on the
amount of reinforcement. If a relatively low percentage of reinforcement is used, the tension
steel begins to yield, causing the widening of the tension cracks in the concrete. Deflection of
the structure becomes noticeable. Where compressive stresses exist (at the top of the beam or
slab), concrete crushing can begin as a secondary mode of failure. Excessive deflections, wide
cracks, and associated spalling provide warning signs of imminent failure. Cracks that open and
close under traffic, diagonal cracks near supports, wide cracks or numerous cracks in any
location should be reported immediately for further evaluation.

If a structure has too much tension reinforcement, the capacity of the concrete in compression is
reached before the reinforcement yields. Concrete crushing occurs in a sudden and brittle
manner, possibly resulting in a catastrophic failure. For this reason, concrete design codes limit
the amount of tension reinforcement to ensure that there is some warning prior to failure.

Axial Stresses

Axial loads are those which are applied along the axis of a member such as a pile or column.
Usually, axial loads have some eccentricity (a load not directly applied along the member’s
central axis) and lateral load, therefore axial loads are normally accompanied by bending loads.
Consequently, concrete members loaded axially perform similarly to those loaded in flexure or
bending and similar failure modes are possible.

Shear Stresses

Shear stresses are caused by loads that tend to slice a member apart in a scissor-like fashion.
Concrete is very poor in its ability to resist shear forces without steel reinforcement in the form
of ties or stirrups. Near the ends of a beam, where high shear forces and low bending moments
are present, web shear cracks may develop, starting diagonally from the bearing area. The most
critical shear area in a concrete beam or slab occurs at approximately a distance “d” from the
face of the bearing area where “d” is the depth of the beam or slab (note that the depth does not
include the ballast curb). It is not uncommon for older, cast-in-place slabs to have minimal shear
reinforcing. Vertical shear reinforcement in areas of high shear, in the form of ties or stirrups, is
used to resist high shear forces. A combination of shear and flexure cracks can be seen in those
areas where shear and moment forces are both significant.

Concrete Deterioration

General

Concrete is used as a part of almost every railroad structure, whether in the foundations, the
abutments and piers, and/or the superstructure. In order to properly inspect concrete structures it
is necessary to be able to identify not only signs of deterioration of concrete, but the structural
behavior of the various portions of the structures. This section is devoted to familiarizing the
inspector with, reminding the inspector of, and enabling the inspector to recognize problems with
concrete structures.

Causes

There are many factors that cause concrete deterioration:

• Freeze/Thaw: Since concrete is porous, it absorbs water. In areas where winter


temperatures are below freezing, the absorbed water freezes and expands creating large
forces internally. This expansion can contribute to many defects in concrete including
cracking, spalling, and scaling. Once concrete is cracked, water will continue to flow
into those cracks and, as it expands with each freeze cycle, it will expand the crack. The
cycle alone can cause serious deterioration of a concrete structure. See Figure 8-1.

• Deicing Salts: Concrete railroad bridge piers and spans are also subject to deterioration
where they cross roadways where deicing salts are used. Salt increases the water
retention of concrete and may chemically attack the structure. Salt is also a concern at
coastal facilities or other locations where brackish water may be found. Salt destroys the
protective alkali coating on rebar in concrete.

• Large Temperature Variations: Large temperature variations, especially sudden swings,


can have adverse effects on concrete structures. Differing temperatures between the
internal mass and the surface of the concrete creates thermal stresses. Large aggregates
can also have a different coefficient of expansion than the cement.

• Unsound Aggregate: As with any material, concrete is only as strong as its weakest
component. If a weak or unsound aggregate was used for the concrete mix, the concrete
may be much weaker than would be anticipated. Many railroad structures were
constructed before modern day quality control was in effect. See Figure 8-2.

• Reactive Aggregates: Certain aggregates are referred to as reactive aggregates, that is,
they have certain properties that react negatively with the cement paste and subsequently
weaken the structure, sometimes dramatically. Problems caused by reactive aggregates
are more predominant in a warm, moist environment. Signs of reactive aggregate include
widespread map cracking and large areas of crumbling concrete with spalls. See Figure
8-3.

• Sulfate Compounds: Certain compounds found in soil and water will, if exposed to the
concrete, cause rapid deterioration of the cement paste and the concrete itself.

• Water Seepage: Water seepage through cracks and voids in concrete dissolves some of
the material and causes corrosion of reinforcement and efflorescence (a white, powdery
deposit) at the surface of the crack where the dissolved material collects. See Figure 8-4.

• Chemical Attack: Many chemicals can attack concrete. Although it is not a common
occurrence for railroad structures, it may be of concern in certain circumstances such as
derailments of tank cars, certain atmospheric pollutants (especially in urban areas), and in
industrial facilities where certain bridge components may be exposed to various
chemicals.

• Abrasion: The most common location for this type of damage to railroad bridges is the
top surface of the deck and is caused by ballast abrading the deck. Obviously, this type
of damage is not normally visible during routine inspections. Abrasion damage can also
occur in waterways that carry large amounts of sand or other abrasive materials,
especially if the stream velocity is high. See Figure 8-5.

• Shrinkage: Concrete shrinks as it cures and goes through the chemical process of
hydration. Although most of this occurs during the initial set, it does continue for a long
time and can cause cracking of concrete, particularly if the concrete is restrained in some
fashion.

• Structural Movements and Damage: As with any structure, the design of concrete
structures or members is based upon certain assumptions about the method of support,
load applications, etc. Should conditions in the field differ or change from those assumed
during design, the structure may be subjected to loads for which it was not designed,
causing cracking or other problems. Common causes of movement on railroad bridges
are foundation settlement, scour and erosion, slope failures, overloads, increased impacts,
unintended fixity, or even the relocation of tracks on the structure. See Figures 8-6 & 8-7.

Surface Indications of Defects

In order to accurately describe concrete defects found during inspection, it is important to


recognize the surface indications of defects, and to utilize standard terminology to describe those
signs of defects. While there may be some slight variations in the description of certain types of
defects, the following terms are widely used within the industry and should be utilized to the
extent possible.

• Scaling: Scaling refers to the loss of mortar or cement between pieces of aggregate. Light
scaling refers to the loss of a small amount of mortar with the surface of the coarse
aggregates exposed. Medium scaling indicates some loss of mortar between the
aggregates. Heavy scaling indicates a significant loss of mortar between coarse
aggregates and severe scaling is used to identify scaling serious enough to cause the
actual loss of coarse aggregates. See Figure 8-8.

• Cracks: On vertical surfaces, cracks are normally described as vertical (up and down),
horizontal (sideways), or diagonal (at an angle along the face). On flat surfaces, cracks
may be described as transverse (perpendicular to the length of the span), longitudinal
(parallel to the length of the span), and diagonal (running at an angle to the span). When a
series of cracks travel in all directions and intersect, usually on a close spacing, it is
referred to as map cracking. Try to note the width and length of significant cracks and, if
possible, probe the depth to determine the extent of damage and as reference for future
inspections. See Figures 8-9 & 8-10.
• Spalling: Refers to a portion of concrete broken out to some depth, usually in a somewhat
circular pattern. There is usually a fracture plane that forms roughly parallel to the
surface of the concrete. Spalling involves the breakout of a section of concrete including
both mortar and aggregate whereas scaling involves the loss of mortar leading to a loss of
aggregate. See Figures 8-11 & 8-12.

• Delaminations or Hollow Areas: Portions of a concrete surface that sound hollow when
struck with a hammer indicate a fracture plane at some depth below the surface of the
concrete. Concrete often has to be chipped out in this instance to determine the depth of
the fracture plane. See Figure 8-13.

• Honeycombs: Hollow spaces or voids that may be present within the concrete.
Honeycombs are caused by improper consolidation during construction, resulting in the
segregation of the coarse aggregates from the fine aggregates and cement paste.

• Popouts: Conical fragments that break out of the surface of the concrete leaving small
holes. Generally, a shattered aggregate particle will be found at the bottom of the hole,
with a part of the fragment still adhering to the small end of the popout cone.

• Efflorescence: Efflorescence is a white deposit on concrete caused by crystallization of


soluble salts (calcium chloride) brought to the surface by moisture in the concrete.

SECTION 2
INSPECTION OF CONCRETE & MASONRY STRUCTURES AND FOUNDATIONS

General

All of the indications of concrete deterioration and defects need to be looked for during all
structural inspections.

Where cracks are found, note the width, orientation, and location. Widths and lengths of
structural cracks should be marked and dated to monitor crack progression. If possible, view
cracks under load to determine if and how much they are opening up.

Location, size, and description of unsound areas, spalling, scaling, or other deterioration should
be noted.

Underwater Inspections

The need and frequency for underwater inspections should be evaluated by an engineer for every
structure having continuously submerged components. These inspections should identify the
channel bottom conditions and presence of any scour, extent of foundation exposure and any
undermining, and all deterioration and damage below water. Consideration should be given to
monitoring for scour by sonar or other means during a high water event. Chapter 8 of the
AREMA Manual for Railway Engineering provides additional guidance on underwater
inspections in the Commentary section.

Foundations

The foundation is the underlying element which supports the substructure unit, whether an
abutment or pier. Most substructure units rest on a foundation of spread footings, driven piles,
drilled foundations, or some combination of these. The substructure loads are transmitted
through the foundation to the supporting strata of soil or rock. As a result, the capacity of the
foundation is directly related to the capacity of the soil or rock upon which it is founded.

Many bridge structures have concrete and masonry incorporated into their foundations,
abutments, and piers. There are many types of construction as well as types of problems or
defects. Some problems are common to all types of foundations and others are unique to a
specific foundation. Different types of construction can have very different responses to the
conditions at any particular structure.

Foundation Types and Problems

• Spread Footings: The footing may not be large enough to carry the loads, resulting in a
failure of the soil underneath the footing. This would normally result in a vertical
settlement of the footing with some rotation also quite likely. Another possible cause of
settlement is the erosion of soil or weathering of stone underneath the footing, reducing
the area of the footing that can carry load, and resulting in some rotation in the direction
of the undermining. Settlement in spread footings can also be caused by a slope failure,
where a wedge of soil containing the footing slips and rotates. This would most
commonly occur on the side of a slope or hill. In addition to likely settlement, a slope
failure would result in some translation, or horizontal movement of the foundation.

• Pile Foundations: There are two types of pile foundations, bearing piles and friction piles.
Bearing piles are piles that are driven down to a very hard layer of soil (hardpan) or rock
and their capacity is based on the bearing at the bottom of the pile. Friction piles are
driven through soil and develop their capacity through the friction developed on the side
of the pile with the surrounding soil. The loss of penetration due to soil erosion, scour, or
excavation around a friction pile can significantly reduce the capacity of that pile or pile
grouping, resulting in settlement or failure of the foundation. That loss of penetration can
also cause stability problems for an abutment or pier as it loses lateral support. Any
apparent loss of pile penetration or changes in condition from the original construction or
previous inspection should be noted on the inspection report for evaluation and
monitoring. Piles are also susceptible to slope failures as described for spread footings.
Friction piles are the most susceptible, since it is possible for an entire pile grouping to
slide in a rotational failure. Bearing piles are less susceptible to slope failures, but can
still be damaged.
• Drilled Piles and Shafts: Drilled pile and/or shaft foundations are similar to pile
foundations, only they are drilled and placed, as opposed to being driven with an impact
or vibratory hammer. Problems to look for in a drilled foundation would be very similar
to a pile foundation.

Inspections for Foundation Problems

Where a structure crosses over a waterway, the inspector should note the condition and
alignment of the waterway. The condition of the slopes and any slope protection (such as riprap)
should be noted along with any indication of debris accumulation. The inspector should note any
indication of damage from marine collision, ice, or debris.

The inspector should note any changes in the alignment of a waterway both upstream and down
stream and the resulting effect that they may have on the structure. A major change in the
alignment of a waterway may place it outside the spans intended to carry the majority of the
flow. A change in alignment may also cause the water to attack the piers at a different angle
which could cause additional drift accumulation or scour.

Where scour is possible, the channel bottom at piers and abutments should be checked by
sounding, probing, or other means.

Foundation Settlement

There are many indicators of foundation settlement, which indicate a problem with the
foundation:

• Horizontal Movement: Spans pushed tightly against the backwall of an abutment


(possibly even breaking the backwall), or spans pulled off of their bearings. See Figure
8-14.

• Settlement: Large vertical cracks in the breastwall or wingwalls. Vertical and/or


horizontal displacement in the track over piers and abutments. See Figure 8-15.

• Rotation: Front faces of abutments out of plumb or rotated from original construction,
and separation from the wingwalls.

• Slope Failure: Open cracks with vertical displacement in the soil near a foundation
element. The problem is often accelerated when a side slope, saturated by high water,
sees a sudden drop in water elevation, leaving a heavy and weakened soil wedge that
could slip or rotate toward the channel bottom.

Undermining

As mentioned earlier, undermining is the erosion of material from under a footing. Undermining
of spread footings is of particular concern due to the reduction of load capacity. The inspector
should note the location and size of the undermined area. Also determine if the undermining
appears to be continuing or active.

Where undermining exposes piles or drilled foundations, note the number of piles exposed and
the horizontal and vertical dimensions of the void.

Undermining of abutments can create an additional concern. As material is washed away from
under the footing, material from above will often fall into that space, potentially creating a large
void behind the abutment. If that void becomes large enough, the track structure above can
collapse suddenly under train loads.

Piles and Drilled Shafts

Alignment and condition of exposed portions of piles and drilled shafts should be recorded.
Impact damage from debris, vessels, or vehicles should also be noted. Piles should be
investigated for soundness. Loss of section and cracking should be noted. Deterioration of piles
may be especially severe in a marine or brackish environment, particularly in the tidal zone.
Note condition of the connections between the piles and cap as well as any bracing.

Exposure to the atmosphere of timber mats or untreated timber piles may lead to rapid
deterioration of the timber. Note where degradation of the channel or other conditions expose
timber that was previously in a continuously wet environment.

Substructure Types and Problems

Most substructure units fall into one of two categories, abutments or piers. Sometimes different
names, such as bents may be used, but for the purposes of this handbook the following
definitions are used:

Abutments and Wingwalls:

Substructure units at the end of the bridge support the bridge spans, and retain the approach
backfill, as well. Abutments may be plain concrete or rubble filled concrete (both unreinforced),
masonry, or reinforced concrete. Plain concrete abutments and rubble filled abutments are
gravity walls and are called that because their own weight is enough to keep them from sliding,
overturning, or failing from backfill and train loads. Gravity walls do not require reinforcement
to work, but may still have temperature steel (steel reinforcement to resist the tensile stresses due
to temperature changes). Cantilever walls are always reinforced and usually much lighter than
gravity walls. They have a large footing to use the weight of the backfill to keep them from
sliding or overturning and they utilize reinforcing to resist lateral loads (described below) as
opposed to just structural weight as a gravity wall does.

In addition to carrying the loads of the bridge and train, abutments must also be capable of
withstanding large lateral loads placed on them from the soil (backfill) as well as the train on top
of that backfill (surcharge). This horizontal force on the back face of an abutment (or retaining
wall) makes the wall want to slide and/or rotate. It also produces very high bearing stresses
along the front edge of the foundation, which can cause differential settlement.

Piers

Piers are intermediate support units. Piers may be concrete, steel, or a combination and can be
very short or very tall and slender. Piers can also be gravity or cantilever and are similar in
function to an abutment except they do not have to resist the backfill loads of the abutment.

Indications of Substructure Problems

Because bridge foundations and the substructure units are so closely interrelated in their
behavior, indications of distress in the substructure elements are frequently the results of, and are
indications of problems with, the underlying foundations. During inspections, always look for
and note any signs or indications that there may be problems with the support capacity of the
bridge foundation as previously noted.

Inspection of Abutments and Piers

When inspecting abutments and wingwalls, look for large vertical cracks which may indicate
differential settlement or rotation. There will often be some offset in the faces of the abutment.
It is quite common to find an offset crack where the wingwall meets the breastwall of the
abutment. Note any indications of settlement, horizontal movement and/or rotation. Spalling
and scaling often occur on the top of walls.

Condition of retained fill, drainage and slope protection at abutments should be inspected. Water
saturated masonry or concrete and extent of efflorescence and rust-staining should be noted.
Check weepholes and drains for proper function.

Sound the concrete with a hammer and note any spalling or hollow concrete on the face of the
walls. Take note of any exposed reinforcing steel and its condition. Any crack in the vicinity of
a bridge bearing should be looked at closely to determine if there is still adequate material for the
bearing. Look to see if there is more than one crack intersecting that would allow a chunk of
concrete to fall away from the bearing area in a wedge type failure. Note any other cracks that
are found.

Pier inspections are very similar in nature to abutment inspections. However, several other areas
of concern are more common with piers than abutments. When piers are located adjacent to
highways, look for impact damage from motor vehicles. Concrete will often be scraped away
and in many cases reinforcing steel is exposed and possibly damaged or severed. If there is
damage to reinforcing steel, note the number of bars damaged and the extent of damage received.

Another item to look for adjacent to roadways is damage from deicing salts. There is often
heavy spalling or scaling of concrete and severe corrosion of reinforcing. See Figure 8-16.
Again, note the extent of reinforcing damage. This problem is particularly common on column
type piers.
Piers in waterways may exhibit abrasion damage from the flowing water and its sediments. Note
any section losses or other damage.

When inspecting an abutment or pier constructed of stone or brick masonry (masonry units),
look for similar conditions as noted previously for concrete. In addition, other items to observe
and note include:
• Missing or fractured masonry units.
• Continuous cracks that extend through masonry joints, fractured masonry units, or a
combination of both.
• Missing mortar from joints.
• Evidence of shifting or movement of masonry units.

Masonry structures are often not solid, but rubble filled. Where stones or bricks are missing or
broken, measure the depth of the damaged area and, in the case of bricks, the number of layers of
brick that are missing.

Retaining Walls

Retaining walls are similar in design and construction to abutments. Retaining walls support
lateral loads from the retained fill as well as train or other surcharge loads that are imposed on
that fill. The closer a retaining wall is to the tracks, the larger the lateral load that the wall must
support. Inspection of retaining walls is very similar to inspection of abutments and wingwalls.
Observe any signs of settlement, movement, or leaning in addition to structural defects. Also
look to see if there are weep holes near the bottom of the wall face, and if they are functioning.

Bridge Seats and Bearing Areas

Bridge Seats: Those areas on tops of the abutments or piers where the bridge bearings are placed.
Generally, bridge seats will be raised slightly above the tops of the abutments and/or piers to
promote drainage away from the bearings.

Bearings: The devices that accept the loads from bridge spans and transmit those loads to the
abutments and/or piers.

Inspection of Bridge Seats and Bearing Areas

Bridge Seats

Carefully inspect the bridge seats to look for problems at the bearing locations. A powdery or
mud-like accumulation around the edge of a bearing indicates that there is some pounding during
the passage of trains that is starting to pulverize the concrete. Sometimes it is obvious that the
bearing has already pounded into the concrete. Many times there will be local crushing or
cracking at the front edge of a bearing due to span deflection during the passage of trains. Also
see if the bearing areas appear to be holding water and if there are cracks anywhere in the bridge
seat which would allow water in, a serious concern in freeze/thaw regions of the country. Check
for loose or missing anchor bolts. See Figures 8-17, 8-18 & 8-19.

Bearing Areas: There will often be concrete blocks or pads on top of the bridge seat underneath
the bridge bearings. Inspect these blocks carefully for cracks, especially if they are shallow (8”
or less). If there is substantial cracking, try to determine if the block is reinforced to prevent any
sudden failure. See Figure 8-20.

Carefully inspect bearings that sit upon mortar-type grout pads. These pads do not hold up well
to railroad loading in the long term, and often crumble due to high impacts. Note bearing repairs
that tend to hold water in or near the bearing areas, especially in areas with freezing
temperatures. See Figures 8-21 & 8-22.

A common problem found at bridge bearing areas is the collection of debris, often consisting of
track ballast. Restricted movement of an expansion bearing can lead to crack development or
even breakout of the concrete around the bearing. Bearings buried in debris should be noted
since this condition can lead to the premature deterioration of the bearing area. Obviously, a
bearing that is covered with debris cannot be properly inspected. Bearings should be cleaned to
the extent possible before or during the inspection. Note all deficiencies observed.

Superstructure

The superstructure consists of the bridge spans which carry the train loads from abutment to pier
and from pier to pier.

There are two basic types of reinforced concrete construction utilized for these spans:

Reinforced Concrete (Cast-in-place and Precast)

Concrete spans are fabricated by having the reinforcing steel placed inside of forms and then
having concrete cast around the reinforcing steel. Cast-in-place reinforced beams or slabs are
cast at their final location. Precast concrete beams are cast at a location other than their final
location, and then picked up and placed in their final locations. Most older spans are reinforced
concrete slabs.

Prestressed/Post-tensioned Concrete

Prestressed/Post-tensioned concrete spans are cast in forms that contain both conventional
reinforcing and wire reinforcing strands. In prestressed concrete beams the strands are tensioned
prior to the placing of the concrete and then the tension is released after the concrete is cast and
has gained enough strength to impart compression into the concrete element. In post-tensioned
concrete beams the wire strands are placed in hollow tubes, the concrete is cast and allowed to
set, and then the strands are tensioned and locked into place by anchoring devices placed at the
ends of the beams.
Inspection of Concrete Superstructures

Conventional reinforced concrete has steel reinforcing bars placed in it at the time it is poured.
For simple spans, as the span is loaded and deflects downward, the bottom of the span goes into
tension, the top compression. Since the concrete is weak in tension, the reinforcing steel carries
the tensile forces. Small cracks in the bottom of conventionally reinforced concrete beams are to
be expected.

Common Defects

The most common defects in reinforced concrete spans are spalling on the underside of the slabs
and spalling or scaling of the ballast curb and corner sections. These items should be noted in
the inspection report. Efflorescence on the bottom of the slab indicates moisture penetration.
Early deterioration of the concrete is often indicated by discoloration. When concrete has been
lost on the underside of the slab, note the depth of spall and if rebar has been exposed. If
significant amounts of rebar are exposed, make a detailed sketch or notes showing how far along
the length of the slab the steel is no longer bonded to the concrete.

Check the sides and bottom of slabs, under load if possible, for cracks that appear to be opening
up. Carefully note any diagonal cracks in the critical shear areas of the slab or beam, note the
size and length and, if possible, see if they move or open under traffic. If excessive span
deflection is noted, it should be reported immediately to the proper authority.

Verify that the ballast curbs are still holding ballast and look for cracks and spalling. Also note
the size and location of any other cracks found.

Many times concrete spans rest directly on the pier or abutment seat with no type of bearing or
pad. Inspect the edge of the bridge seat for cracked or breaking concrete. Also inspect the span
itself for crumbling or spalling at the edge of the seat.

Prestressed Concrete

Prestressed concrete presents a different set of inspection guidelines. Prestressed and post-
tensioned concrete differ from conventional reinforced concrete in that very high strength steel
wire or strands are pulled to a very high tensile stress. In the case of prestressed concrete, these
wires are pulled tight and the concrete for the beam or slab is placed. The wires stay pulled tight
until the concrete cures and bonds to the steel. Then the wires are released. Since the wires now
want to shorten back to their natural length, they pull the concrete tightly together at the bottom
of the beam where the wires are located. The end result is that when the span is placed in the
bridge, and a train travels across it, all of the concrete remains in compression, even on the
bottom.

Post-tensioned concrete behaves very similarly to prestressed concrete. The difference is that the
concrete is placed with voids for the post-tensioning strands. Once the concrete has cured, the
strands are placed in the voids and then tensioned at the ends with thrust blocks or anchors. The
net result is similar, with all the concrete in compression under service loads.
Because the concrete in prestressed beams is in compression, there should not be cracks in the
sides or bottom of the beams or boxes. The presence of cracks indicates that there has been
damage during shipping or placement, some strand slippage or a combination of these. It is
always a good idea to carefully inspect beams at the time of placement to look for damage during
shipping and handling. Until they are fully in place, prestressed beams are very susceptible to
damage, especially if they receive a lateral load for which they were not designed. Hairline
cracks in the top of prestressed beams are generally due to shrinkage of the concrete and not of
major concern (an exception to this would be cracks found in the top of continuous spans in the
support, or pier area). Chamfers of boxes should be inspected for cracking which may extend
along the sides or bottom of the girders.

Inspect the ends of prestressed beams for evidence of strand slippage. The end of the strand
would be recessed into the concrete (condition not always visible if recess is filled with grout).
Also look for cracks near the end of the beam. These could be “bursting cracks” which result
from very high compressive stresses at the ends of the beam. In some cases, not enough
reinforcing steel has been added at the ends of the beams to withstand these stresses. See Figure
8-23.

Carefully inspect any prestressed beam that has been struck by a high load on a vehicle passing
under it. Look for any cracks or indications of strand slippage. If strands are exposed, note how
many and for what length they are exposed.

Arches

When inspecting arches, note whether it is constructed of concrete, brick masonry, or stone
masonry as well as the overall condition of the material. Check the foundations for settlement,
shifting, scour, or undermining as previously discussed.

Check the arch ribs and bearing areas at the springlines for any loss of cross section due to
spalling or cracking. Open spandrel columns and walls should be inspected with particular
attention to areas near the interface with the arch rib and cap. Arch ribs connected with struts
should be inspected for diagonal cracking due to torsional shear (twisting).

When inspecting arches, look for uniform lines and surfaces. Note any bulges, voids, or other
instances where the shape of the arch is not true and consistent. Look for changes in horizontal
alignment, sags in the arch crown, bulges in the sidewalls, longitudinal cracks, or expansion joint
failures which may be signs of settlement or overload.

Masonry was often used for arch construction. Note any specific defects as discussed under the
inspection of masonry abutments and piers.

Fire Damage

Fire can cause considerable damage to concrete structures. The inspector may or may not be
alerted to the fire when it occurs and should be able to identify the telltale signs of fire damage
including discoloration, spalling, and cracking. The extreme heat will reduce the load carrying
capacity of the structure, especially when prestressing steel is rendered ineffective through
relaxation. The inspector may not be able to make an accurate assessment to the degree of
damage and should notify a qualified engineer to make the final judgment.

CONCRETE BRIDGE INSPECTION CHECKLIST

Track
• Track alignment and surface profile. Determine if any defects are track or structure
related.
• Condition of rail, fasteners, and other track material (OTM).
• Condition of track or bridge ties.
• Ballast.
o Clean or fouled.
o Depth.
o Adequate shoulder.
o Ballast retainers.
o Locations that are missing ballast.

Walkways
• Walking surface.
• Posts and handrails.
• Mounting brackets.
• Hardware/fasteners.

Superstructure
• Overall alignment – Vertical and horizontal.
• Impact damage.
• Cracking.
• Spalling or scaling.
• Exposed reinforcing.
• Ballast leaks.

Bridge Seats/Bearing Areas


• Gaps, or lack thereof, between bearing components and between spans and abutment
parapets.
• Bearings pounding on seats.
• Cracks in bearing areas.
• Potential for ponding of water around bearings.
• Loose or missing anchor bolts.

Substructure, Retaining Walls, and Foundations


• Cracking.
• Spalling or scaling.
• Exposed reinforcing.
• Undermining.
• Scour or erosion.
• Settlement, rotation, or lateral movement.
• Evidence of slope failures.

Masonry
• Missing or fractured stones or bricks.
• Continuous cracks.
• Condition of mortar joints.
• Shifting or movement of masonry units.
• Depth of damaged areas.

Emergency

If the inspector discovers a bridge condition that affects the integrity of the
bridge under train loads, contact the railroad dispatcher and responsible
authority to stop any trains and arrange for immediate repairs. Criteria for
the limits of acceptable/tolerable conditions must be provided by the Railroad
Owner or their designated engineer based on the principles of acceptable
rating contained in the appropriate chapter of the AREMA Manual for
Railway Engineering.
Concrete & Masonry Bridges and Foundations

Figure 8-1 Cracks in Breastwall Figure 8-2 Severely Deteriorated Concrete

Figure 8-3 Map Cracking Due to Reactive Figure 8-4 Efflorescence


Aggregate

Figure 8-5 Concrete Abrasion Figure 8-6 Structural Cracks


Figure 8-7 Wingwall to Backwall Crack Figure 8-8 Scaling

Figure 8-9 Crack at Bearing Figure 8-10 Map Cracking

Figure 8-11 Spalling Figure 8-12 Spalling Under Bearing


Figure 8-13 Hollow Concrete Figure 8-14 Backwall Tight Against Span

Figure 8-15 Foundation Settlement – Note Bearing Figure 8-16 Deicing Salt Damage
Shims

Figure 8-17 Broken Stone at Bearing Figure 8-18 Concrete Breaking at Front Edge of
Bearing
Figure 8-19 Concrete Cracking Under Bearing Figure 8-20 Cracked Raising Block

Figure 8-21 Failed Grout Pad Figure 8-22 Pumping Bearings

Figure 8-23 Bursting Cracks Figure 8-24 Shear Compression Failure


Figure 8-25 Shear Tension Failure
CHAPTER 9
STEEL BRIDGES

SECTION 1
IRON AND STEEL BRIDGES

General

In the inspection of Iron and Steel Bridges, the inspector is primarily looking for physical defects in
the steel such as deterioration of the metal (corrosion losses), cracks in critical locations, damaged
members, broken and/or missing fasteners and how the bridge responds to the passage of trains.

This chapter is formatted to first discuss basic information the inspector needs concerning the
behavior of steel as a material and its usage in bridges. This includes fatigue of the metal, fracture
critical members and fabrication. Thereafter, common steel bridge items such as fasteners, bearings,
and floor systems are each discussed followed by bridge spans from short beam spans to girder spans
to truss spans.

History

Cast iron and wrought iron were used in early timber bridges in joint connections and used alone or
with other members such as vertical and main chord members. Wrought iron bridges were
constructed primarily in the 1840's and 1850's, but there remains in use a fair number of bridges
from the 1880's that have wrought iron components. Such wrought iron components can be
identified by a spark test as wrought iron has very distinctive sparks from that of steel, see Figure 9-
1, this Chapter. Also, wrought iron where observed in a somewhat deteriorated condition has the
appearance of “strings” or series of longitudinal lines in the metalwork surface.

Steel production in sizeable quantities for bridge usage was developed by the 1870's. There are steel
bridges from the 1870's that remain in service today. The “golden age” of steel bridge construction
occurred from the 1890's through the 1920's. Although the life expectancy of steel bridges was once
thought to be 100 years, a review of steel bridge ages listed by major railroads and short lines
indicate that bridge life exceeds 100 years and with care, railroads are further extending bridge life
through retrofit/rehabilitation programs. In the case of earlier bridges retired from service, generally
the reason is for lack of carrying capacity for present traffic needs, wear of components or corrosion
deterioration.
Figure 9-1 Spark Test for Sorting Metals
Steel - Material

Not all steel is alike. Steels can be alloyed to satisfy various needs such as high strength to minimize
the weight of bridge members, toughness for miter rails and fracture critical members, weathering
characteristics for various environments, and for other requirements. Steels for bridge building have
been chosen based on the following properties: tensile strength, ductility, fatigue strength,
toughness, and weldability. Steels presently used in the design of bridges can be reviewed in the
AREMA Manual for Railway Engineering. Field identification of a particular steel of any age can be
done through the removal of samples (coupons) for determining strength and chemical composition.
The removal location of coupons from a bridge must be determined by an engineer with knowledge
of where to remove material without affecting the strength of the member.

Steel - Fabricated Members

Unlike concrete that is molded or wood that is cut into rectangular shapes, steel is rolled into plates
and shapes to common industry standards. Earlier bridges often had shapes that were only fabricated
by a particular mill. Given the age of the bridge, the field measurements of a shape and the
identification of the mill (stamped in the shape), the inspector can be directed to books that identify
that shape. The American Institute of Steel Construction (AISC) publishes the Manual of Steel
Construction which lists all currently rolled shapes and plates, giving dimensions and weights of
each piece. The basic rolled shapes used in bridge construction include: wide-flanged beams,
standard beams, H-shape (bearing piles), channels, angles and T’s.

The AISC has also published a comprehensive list of historical shapes in the Rehabilitation and
Retrofit Guide. Designers use shapes and plates in combination to form various fabricated
configurations to achieve the required strength for tension, compression and/or bending members.

Tension Members - Tension members are susceptible to crack initiation and development
from fatigue and flaws that cause stress concentrations in the member. Flaws may include rolling
inclusions or laminations, discontinuities such as tack welds or intermittent welds, and notches from
accidental scrape marks, errant weld strikes or from corrosion pitting and sharp edges. Additional
information on the sensitivity of tension members to fatigue and fracture critical details follows in
this Section.

Tension members may be as simple as a rod or what is known as an eyebar. These members only
have strength in tension and will buckle in compression. Rods used for bracing are often in a crossed
pattern. Depending on the direction of a force, one rod will be in tension and the other rod will not
carry load. When the load is reversed the opposite rod will take the tension load. Rod ends must be
carefully inspected as many of the earlier rod ends were “loop” rod ends; where the rod end was
forged into an eye.

Rods
In earlier years, eyebar “heads” were formed by heating and upsetting the bar material until enough
material was available to make a head. This accounts for many old eyebars having rough head edges.
Eyebars presently are fabricated from a single plate cut to the familiar eyebar shape. The pin holes in
a set of eyebars are commonly drilled as a group.

Eyebars

Other common tension members include rolled I-beams, a pair of channels with lacing bars that form
a box shape, vertical web plates with corner angles with top and bottom lacing bars and batten plates
to help hold the shape, solid web and cover plates welded to form a box, and many other
configurations.

Other Tension Configurations

Compression Members - Compression members are susceptible to buckling and therefore


must be rigid or adequately braced. Many compression members can be readily identified by their
fabricated shapes such as: a pair of channels with legs facing outward with a solid top cover plate
and bottom lacing bars (top chords and end posts), a pair of channels with legs facing inward with
top and bottom lacing or batten plates (diagonals), a pair of channels with legs facing inward with
lacing on both sides (posts), current fabrication commonly is for four plates welded into a box.

Compression members are not fracture critical nor fatigue sensitive.


Compression Member Shapes

Bending Members - Bending members are generally beam shaped consisting of a top and
bottom flange and a solid web plate, resembling an I. Smaller bending members may be rolled
beams such as wide flange beams used for short spans. For longer spans and heavier loads, the
member may be a web plate with top and bottom flange angles with cover plates. Girders are large
bending members with deep web plates that require stiffening to keep it from buckling and top and
bottom flange angles with cover plates. Many girders currently fabricated are either of all welded
plates or may have flanges welded to the web with bolted web stiffeners.

Bending Members

Steel - Fatigue

Fatigue is a process that causes accumulating damage and eventually failure of metal subjected to
cyclic loading, i.e., generally from the passage of railcars on a bridge. With the passage of railcars,
some members receive a load cycle for every set of wheels crossing while other members only
receive one cycle per car or one cycle per train, see Loading Diagrams in Truss Section, this
Chapter. Fatigue cracks are caused by primary stress cycles and by secondary out-of-plane distortion
induced stresses. A primary stress crack is formed perpendicular to the direction of force. An
example: a truss hanger may develop a horizontal fatigue crack, generally across the bottom row of
fasteners in the member upper connection. See Figure 9-2. A primary fatigue crack tends to
propagate quickly and if not found and repaired can cause sudden failure. The primary factors
leading to the development of such cracks are the frequency of applied loads, the magnitude of stress
(heavy load applied over a given area), the age (stress cycles) and condition (corrosion) of the bridge
and type of steel, the quality and fatigue sensitivity of fabrication details, and material fracture
toughness. Generally, the stress fluctuation, frequency, and type of detail are the most important
factors. Welded structures tend to be less forgiving of small weld discontinuities than riveted
structures. These welds are more sensitive to repeated stresses and once cracking starts to develop,
it can destroy the member as a result of the continuous path provided by the welded connections.
Riveted or bolted construction generally utilizes multiple plates or shape components and when one
component cracks, the crack does not directly enter into the other components. This is also known as
internal redundancy.

Critical cracks will generally only exist in tension areas. When details are located in compression
stress regions and no possibility of stress reversal exists, fatigue will not occur. Under these
conditions, any crack growth will be confined to a residual tensile zone unless out-of-plane
deformations occur. The usual and most reliable sign of fatigue tensile cracks are the oxide or rust
stains that develop after the paint film has cracked. Experience has shown that cracks have generally
propagated in depth to between one-fourth and one-half the plate thickness before the paint film is
broken. Once a fatigue crack is found, other similar details at similar locations on the bridge should
be checked. This may require the removal of the paint to detect the crack. Non-destructive testing
means such as dye-penetrate, magnetic-particle or ultrasound should be used to find these other
potential cracks.

Secondary fatigue cracks such as out-of-plane cracks are as the name implies, the working of the
metalwork similar to when one bends a paperclip back-and-forth until it fractures. A prime example
of out-of-plane fractures occurs in thin top flanges of stringers and/or girders where ties are in direct
contact and as ties deflect, the flange is bent out-of-plane. See Figure 9-3.

Those members most likely to develop fatigue cracks are those with the larger stress ranges such as:
stringers, floorbeams, and hangers. Influencing the development of cracks are: notches, corrosion,
tack welds, copes, weld terminations and intersecting welds. Weld details most likely to develop
cracks are located at: ends of cover plates, gusset plate welds to transverse stiffeners, groove welds
in flange transitions, and flanges or plates that frame into or pass through webs.

Various details will have different fatigue susceptibility associated with them. The bridge designer
takes these details into account. The bridge inspector should be familiar with those details most
susceptible to fatigue propagation. These details are known as Category D, E and E’ and may be
seen along with other category details in the AREMA Manual for Railway Engineering. A selection
of some critical details is shown in Figure 9-4.

The fatigue failure process consists of three phases: crack initiation, crack propagation (growth), and
failure. The crack first initiates from points of stress concentrations in structural details. Stress
concentrations can result from flaws, geometric details, or out-of-plane distortion. Once the fatigue
crack has initiated, applied cyclic stresses propagate the crack across the section of the member until
it reaches a critical size at which time the member fractures. It has been stated that once a crack is
visible, typically 85% of the energy required to initiate the crack has been expended and that it will
take only 15% more energy to drive the crack to member failure.
There are two common fractures: brittle and ductile. A brittle fracture occurs without warning,
without any deformation and at average stresses below those of general yield. The crack appears in a
long smooth curve. Ductile fractures have plastic deformation resulting in distortion of the member
which provides some visual warning of the impending failure.

Figure 9-2 Primary Fatigue Figure 9-3 Secondary Fatigue Crack -


Crack - Hanger, Primary Stress Crack Out-of-Plane Bending
(Perpendicular to Direction of
Force)

Figure 9-4 Fatigue Details Category D, E and E’


Steel - Fracture Critical Bridge Members

A fracture critical member (FCM) is a member in tension, or with a tension element, whose failure
would probably cause a portion of, or the entire bridge to collapse. When inspecting steel bridges,
the inspector must be able to identify an FCM by sight or based on previous reports and drawings.
Fatigue is the primary cause of failure in fracture critical members.

Redundancy is important as it refers to the load distribution to members and member internal
makeup. Redundancy, as it refers to load distribution, relates to alternate means to carry the load
should a member or element fail. A 4-stringer floor system is redundant, a 2-stringer system would
be fracture critical. A 4-eyebar bottom chord is redundant, a 2-eyebar chord is non-redundant.

Internal member redundancy exists when a member contains several elements which are
mechanically fastened together so that multiple load paths are formed. Failure of one member
element would not cause total failure of the member. A riveted stringer consisting of a web plate and
four flange angles is internally redundant. Should a crack develop in one flange angle, it will not
travel into the web or adjacent flange angle. In a stringer consisting of a web plate and welded
flanges, a crack developed in the flange will travel across the flange and into the web until failure.
Thus, the critical nature of inspecting welded members and the importance of fracture critical
member inspection for welded members.
Steel - Deterioration

Steel deteriorates primarily from corrosion of the material, but also can be rendered unusable as a
result of fatigue, overload, collision and fire.

Corrosion - is actually the transformation of the material to its oxide form from a reaction of
oxygen, water, or other agents. Corrosion is an electrochemical process. It requires metal, an
electrolyte and current flow. It occurs between metal areas having a higher tendency to corrode
(anode) and metal areas having a lower tendency to corrode (cathode). An electrolyte which allows
current flow must be in contact with the anode and cathode for corrosion to occur. Usually this
electrolyte is water. The corrosion product for steel is iron oxide, better known as rust. The rate of
deterioration depends on the type of steel, surface protection, stress, the presence of pollutants,
debris, other factors.

It is obvious that a bridge with unprotected steel in an area with high temperatures and high moisture
available will corrode much faster than a similar bridge in a desert. Studies have shown that
corrosion rates are approximately 2.75 times higher when salt is present than when it is not.

A-7 steel that was produced from 1901 to the 1960’s was the standard of that time. The A-7
specification was withdrawn in 1967. Nearly all of the bridges built during that time period were
built of A-7 steel or, during the part of the period before 1935, of steel with similar mechanical
properties and chemistry specified in the AREA Specifications for Steel Bridges. In the 1935 and
later specifications, A-7 steel was specified. That steel requires coating for protection.

A-36 steel was introduced in 1960 and remained a common bridge steel until recently. It also
requires coating protection.

A-94 known as “silicon steel” was used in the 1930's and 1940's and has slightly more
corrosion resistance than the A-7 and A-36 steels.

A-572 steel has manganese, and columbium or vanadium alloys and as such offers about
twice the atmospheric corrosion resistance as A-36 steel.

A-588 steel is known as a “weathering” steel. It has alloys of nickel, copper and chromium
which give it about four times the atmospheric protection as A-36 steel. When unpainted it forms a
patina or tightly adhering oxide which becomes the protection. It is important that this patina
develops and that corrosion product such as sheets of rust do not develop.

A-709 structural steel for bridges, introduced in 1974, includes grades 36, 50 and 50W,
equivalent to A-36, A-572 and A-588, and is the steel currently specified for bridges.

The affects of corrosion are the general loss of metal which reduces the member section and
diminishes the member’s carrying capacity. It can also affect the fatigue life of the member because
of the increased stress range. Corrosion can also cause stress raisers with the formation of holes and
rough edges which increases stress concentrations and often becomes the location for the initiation
of cracks. With advanced corrosion, the formation of significant corrosion product can “freeze”
moving parts, such as expansion devices and hangers and cause the structure to behave differently
than intended. The corrosion product can also cause unintended movement of components. Pack rust
can generate pressures up to 10,000 psi. This pressure is sufficient to cause bending of members,
break rivets, push eyebars off pins, crack piers, and collapse bearings. The local bending of plates in
built-up members due to pack rust between rivets can significantly reduce the strength of
compression members.

Corrosion Mitigation - Corrosion of bridge members can be accelerated by environmental


factors such as: nearby industry (power generation, chemical refining, pulp and paper production,
fertilizer manufacture, etc.), salt laden prevailing winds, proximity of the bridge to water, poor
drainage details on the bridge, and pockets of dirt and debris.

Mitigation of corrosion can be as simple as keeping a bridge clean of debris, providing drainage
holes to relieve pockets of water, and by providing paint protection. The painting of bridges is often
dictated or justified for safety, aesthetics, and economic reasons. Painting may be performed in
spots, zones or the entire bridge.

Special bridge inspections may be required to evaluate the extent of corrosion on a bridge to
determine the need for paint protection. A special bridge inspection may be required to evaluate
existing paint protection for suitability to overcoat, the need to replace or to analyze for the presence
of lead, cadmium, chromium, and inorganic arsenic.

Collision - Steel may deteriorate from repetitive collisions and subsequent straightenings
such as mechanical heat straightening the metal. Prime areas of bridges for excessive deterioration
from collisions include the end posts of thru-trusses from wide loads, girder and beam span
underpasses where vehicles have struck the bottom flanges, and from a variety of other impacts
(drift, marine vessels, ice) to the bottoms of spans.

Often it is necessary to perform emergency field repairs to stabilize and repair collision damage.
When it is necessary to perform emergency field welding, a quick spark test in the field can be used
to determine if the material is wrought iron or steel, and whether or not the steel is weldable. Using
an emery wheel and touching the steel will throw sparks. Those sparks are actually bursts of carbon
exploding. Thus one can readily determine if the carbon content is low enough to allow for
emergency welding.

Fire - All railroads have experienced bridge fires, some being more dramatic than others, but
all requiring inspection and proper evaluation. Of interest in a fire evaluation is the type of
combustible material involved, the temperature experienced by the various components of the bridge
and for what duration of time were they exposed. Temperature and duration relate directly to the
extent of damage. There are usually some indicators at the bridge site that will assist in determining
the temperature sustained during the fire. Several common materials and their approximate melting
points are:
Plastic water bottle 300 - 450 F

Glass bottle 750 - 900 F


Aluminum cans 1200 F
Copper (rail jt. continuity wire) 2000 F
Babbitt (bearing lining) 600 F

Color can also be an indicator such as observed in the heating of steel:

Red heat, in the dark 750 F


Red heat, in daylight 1000 F
Dark red 1300 F
Dull, cherry red 1500 F
Bright, cherry red 1800 F

The duration of the fire is generally established by someone reporting or observing the fire.

Steel bridges generally behave rather well in a fire environment; however, properties are altered
while heated and if heated sufficiently, the steel properties can be permanently altered. At a
temperature of 1000 F, the strength of steel is about half that of its normal strength and the rate of
expansion is about 125% of normal expansion. At higher temperatures, such as that noted by a
cherry red color, a transformation temperature is reached for some steels and the grain structure of
the steel can be permanently altered. This is particularly of concern for heat treated and tempered
steels.

Besides permanent alteration of steel mechanical properties, probably the most widely experienced
problem is warpage (distortion) and buckling. If the whole structure could be uniformly heated there
would not be any problem; but as normally occurs in a fire situation, there are local areas of high
heat which causes accelerated expansion to occur in some members while nearby members remain
cooler and act as restraints. Thus, for lighter members, buckling and warpage readily occurs.
Distortion also occurs from relieving rolled-in residual stresses in beams and/or from weldments. At
a temperature as low as 600 F, both rivets and bolts may lose their clamping force.

Obviously no traffic should run on a bridge while its temperature is elevated. When the temperature
falls sufficiently to inspect the bridge, the steel will have re-attained its strength. Applying water on
the hot steel may result in increased distortion and embrittlement of the steel.

The inspection should include looking for indicators that may show the extent of heat applied to the
bridge. The occurrences of heavy black scale or pitting are strong signals of possible permanent
damage. The inspection of fasteners is of primary interest and the inspection should extend well
beyond the immediate damage area as sheared fasteners have been known to occur at distant
locations. Primary members with warpage and distortion need to be evaluated to determine whether
they are stable and can adequately carry load. Secondary members, such as bracing can generally
remain in place temporarily and may remain in place permanently, upon review. Often in intense
fires, small bridges with light sections will deflect (permanent sag).

Temporary shoring may be sufficient to restore traffic. It may be quicker and simpler to replace in-
kind small bridges composed of rolled beams or plate girders, than to do time consuming testing and
repairs. For major bridges which cannot be replaced, destructive and non-destructive testing is in
order.

Iron and Steel Bridge Inspection Checklist

Checklists for the inspection of iron and steel bridges are included in each individual section of this
Chapter.

Emergency

If the inspector discovers a bridge condition that affects the integrity of the bridge under train loads,
contact the railroad dispatcher and responsible authority to stop any trains and arrange for immediate
repairs.

Criteria for the limits of acceptable/tolerable conditions must be provided by the Railroad Owner or
their designated engineer based on the principles of acceptable rating contained in the appropriate
chapter of the AREMA Manual for Railway Engineering.
SECTION 2
STEEL - FASTENINGS

General

In the design and construction of bridges, the type, number and spacing of fasteners for the
fabrication and connection of bridge component members is significant. For the fabrication of
members: rivets, bolts and welds are used. For the connection of bridge component members: rivets,
bolts and pins are used. Welding is generally not permitted in the field.

The spacing of fasteners, both in the design of members and in connections, follows standard
fabrication practices and recommended practices of the AREMA Manual for Railway Engineering
and the AISC Manual of Steel Construction. Spacing practices are necessary to assure fasteners are
not spaced too far apart to get spreading between pieces being joined or too close together to cause a
weak plane for tearing. Standard practices give specific “gage” lines to position fastener holes
correctly in rolled shaped. This involves minimum spacing from the edges of rolled and cut edges of
shapes and plates as well as hole positions so that fasteners can be installed without interfering with
each other. In the field, the inspector often will readily note deficiencies as a result of improperly
specified fastener spacing, particularly where spacings are large and member components are
separating from pack rust. In the design of members with multiple plates, such as box members with
multiple web plates, it is necessary to “stitch” those plates together with fasteners for the plates to
act as a unit. The AREMA Manual for Railway Engineering also gives requirements for stitching.

Fasteners in direct tension (pulling on the head or nut) should be closely inspected.

Rivets - Rivets were used in bridge construction from the start of iron and steel bridges until the late
1950's, early 1960's. For a period of time in the 1960's, bridge members were shop riveted and field
bolted. Thereafter the process became a cost prohibitive method of fastening. Rivets are
manufactured with a head and a blank shank. That shank length, like bolts, must satisfy the
requirement of the thickness of the joining pieces “grip length” and the amount of shank material
necessary to form a head. Rivets are heated to a specified temperature, positioned in the hole, held in
position while a pneumatic rivet gun with a cupped head impacts the hot rivet, driving the rivet
shank until it fills the hole and forms a head. The completed rivet upon cooling provides clamping
force on the connected plates. With the rivets completely filling the hole, the rivets act in shear by
transferring loads from one component to the other.

Rivets are checked by “ringing” the rivet. Ringing is a process of striking the rivet head and feeling
the backside of the head to determine if the rivet is tight in the hole. If not, the rivet is loose and
should be replaced. It is unlikely that the inspector today will still find originally installed rivet
defects other than rivets with too small, too large, or poorly-formed heads. The inspector will find
defective rivets now from worn or corroded heads or heads “popped” off from prying action. In
general, if the rivet is tight in the hole and is keeping the plies of steel together as a unit, head loss
can be accepted. However, judgment must prevail in observing whether it is possible for an outer
plate to detach from overall lack of fastener heads.
For a period of time, many rivet heads deteriorated from corrosion (blossomed heads) from
refrigerator car brine water droppings. Many bridge floorbeams exhibit rivet with head loss on one
side where the brine water splashed, yet the rivet heads on the other side of the member are in good
condition.

Rebuilding rivet heads with weldment and weldment used to seal spreading seams are not good
practices. The inspector should closely watch these areas for the development of cracks.

Loose and defective rivets should be replaced with high strength bolts.

Bolts - High strength bolts (ASTM A-325 or A-490) have been in use since the retirement of
riveting. In the design of bolted connections, the friction between the surfaces of joining pieces
(faying surfaces) is considered, and such connections are dependent on all of the bolts in that
connection being properly tightened to develop that friction.

High strength bolts, like rivets, have a manufactured head and shank. A-325 and A-490 bolts
recommended by the AREMA Manual for Railway Engineering, are intended to have a plain shank
in the “grip” distance and the end threaded for the washer and single nut. It is intended to have 1 to 2
threads project beyond the tightened nut. High strength bolts are manufactured as a “black” bolt (no
corrosion protection) or in weathering steel which matches the materials being joined in weathering
characteristics. There are two types of galvanized high strength bolts, the hot dipped galvanized and
the mechanically galvanized. In either case these bolts should not be used in contact with bare steel
(galvanizing is sacrificed). The hot dipped bolts are subject to hydrogen embrittlement which can
lead to fracture; and, the threads on the bolt and nut must be re-chased by the manufacturer after the
parts have been dipped to remove burrs. This can lead to having bolts with undersized threads and
subject to stripping when torqued. The mechanically galvanized bolt has the zinc coating applied
from tumbling in drums with charged zinc pellets. There is no need to re-chase the threads and
hydrogen embrittlement is not a concern.

The A-325 high strength bolt is intended to have a single hardened washer under the turned element
and a heavy hex nut. A-490 bolts are intended to have a hardened washer under both the head and
nut. Inspectors may note that some bolts have a smooth cylinder shaped nut which is a Huck bolt, or
observe bolts with a small sheared tip at the end of the bolt. These types of bolts were hydraulically
tightened. There are a number of other types of high strength bolts including bolts with a rivet head.
All high strength bolts should have identification on the bolt head, giving the manufacturer’s name
or initials, show the numbers 325 or 490 and/or have 3 or 6 radial lines which indicate an A-325 or
A-490 bolt respectively. The use of SAE Grade 5 or 8 bolts which are approximately equivalent to
A-325 and A-490 bolts respectively is not encouraged, as the full length of the bolt is threaded
which results in a slightly undersized bolt shank.

The tack welding of nuts in the field is an improper means to assure that the bolt does not loosen as
they can lead to the development of cracks.

Inspectors should check that bolts are tight and that there is no significant corrosion loss of the bolt
head or nut. Unlike rivets that stay in the hole even when the head is lost, the loss of a bolt head or
nut (depending on bolt orientation) will result in the fastener falling out of the hole.
Welds - A major portion of new bridge components are sections fabricated by weldment of plates
and shapes. The welding processes used in fabrication are per the AREMA Manual for Railway
Engineering which is based on the American Welding Society (AWS) D1.5 Bridge Welding Code.
These welds have been visually and NDT inspected in the shop before delivered to the field.

In new construction, the termination of welds such as at the ends of horizontal stiffeners and on short
vertical stiffeners and for other attachments are locations subject to cracking. Many of the cracks
found in weldments are associated with fatigue induced conditions. Examples of some of the
potential locations for weld fractures are shown throughout this Handbook.

Cracks in welds found in the field are generally from earlier field welding placed under non-
controlled methods and conditions. It is not uncommon to find stringers with field added cover
plates having longitudinal fillet cracks and cracks at the ends of the cover plates or at butt welds due
to the omission of adequate heat during welding.

Tack welds on bridge members should be carefully watched for possible origination of cracks. Tack
welds on fracture critical members should be removed by grinding.

Inspectors should closely observe locations where welds augment bolts in the same connection.
There is a potential for weld cracking at such locations.

Suspicious areas in welds may be checked for cracks by non-destructive testing methods such as
ultrasonic, magnetic particle or dye penetrant methods.

Emergency

If the inspector discovers a bridge condition that affects the integrity of the bridge under train loads,
contact the railroad dispatcher and responsible authority to stop any trains and arrange for immediate
repairs.

Criteria for the limits of acceptable/tolerable conditions must be provided by the Railroad Owner or
their designated engineer based on the principles of acceptable rating contained in the appropriate
chapter of the AREMA Manual for Railway Engineering.
Rivets

Figure 9-5 Head loss due to salt environment Figure 9-6 Floorbeam web – losses from Brine car
drippings

Figure 9-7 Loose rivets Figure 9-8 Corrosion/pack rust formed between
widely space rivets

Figure 9-9 Rivets Figure 9-10 Stretched rivet


Loose and movement between plates
Bolts

Figure 9-11 Deteriorated A-325 bolts Figure 9-12 A-325 bolts


Deterioration due to salt water

Figure 9-13 Connection with loose and missing Figure 9-14 Improper tightening of bolts lead to
bolts loose and missing bolts

Figure 9-15 Nut corrosion due to poor Figure 9-16 Connections with rivets replaced by
environmental conditions – power plant nearby Huck bolts
Welds

Figure 9-17 Welding mixed with rivets results – Figure 9-18 Girder cover plate with weld broken
cracked weld longitudinal
and butt weld

Figure 9-19 Improper welding to eyebars Figure 9-20 Improper welding to hanger

Figure 9-21 Welded rivet heads and crack in the Figure 9-22 Floorbeam patch plate with broken
butt weld of hanger connection welds
SECTION 3
BEARINGS

General

The type of bearing for a span is generally governed by:

• length of span and the magnitude of load to be transferred to the supporting element.
• thermal movement.
• the need to accommodate span deflection.
• the designer’s or owner’s choice.

The overall appearance of the bearings for a span will be somewhat similar, i.e., if flat plates are
used at one end, flat plats will be used at the other end, with the only difference being for the
accommodation of thermal movement. If the expansion end bearings for a span have castings with a
center pin connection, the fixed bearing for that span will have castings with a center pin connection;
however, the expansion bearings will accommodate movement and the fixed bearing will be
stationary.

For a simple span, bearings at one end will be fixed or stationary bearings with expansion
accommodating bearings at the other end. Some exceptions may be noted on short spans such as
those atop towers where both ends are fixed. For larger, multiple span bridges, often the spans are
arranged on the piers to accommodate the fixed bearings for two spans followed by a pier
accommodating two expansion bearings for two spans. This arrangement allows for the reduction of
bending stresses in the pier.

The bearings must be able to resist longitudinal braking and traction forces, wind, and other forces
besides the dead load and live load of the span, thus the need to have adequate attachment of
bearings to the supporting element. Fixed bearings must be able to resist longitudinal forces while
the expansion bearings must be able to resist transverse forces. Some bearings have no inherent
resistance to lateral movement and require separate lateral restraint to transmit lateral forces from the
superstructure to the foundation. Similarly, some bearings have no inherent resistance to uplift and
require separate hold down devices to prevent uplift or separation during earthquakes.

Bearing base size is dictated by the bearing pressure imposed on the supporting material.

Rarely does an inspector find a bearing imbedding into a granite block, but often the inspector will
find that the granite block under the bearing is loose among adjacent stone blocks. Frequently,
bearings on older bridges are noted as pounded into the pier top from an over pressure on the
surface, primarily when the supporting base is on sandstone, limestone or a concrete surface. Once a
bearing starts to embed into the surface, the process accelerates due to water collecting in the
depression and the pulverized fines expelled with the pounding from passing wheel loads.

Surfaces for the support of bearings must be flat and without any irregularities that would cause hard
spots and cause cracking of the bearing casting. Bedding for bearings have ranged from lead sheets
to red lead paste between layers of cotton duck cloth to steel grillages finely machined steel
grillages.

Bearings - Materials

Bearings may be as simple as sliding plates or as complicated as spherical, machined bearings.


Commonly, rollers are cast steel ASTM A-27. The upper and lower elements of older bearings are
castings of A-27 or A-148 material. Many of the newer bearing elements are weldments. The
bearing pins are usually A-668 material. Where bronze sliding surfaces are employed, the bronze
material is ASTM B-22 using one of three possible alloys. Often the sliding bronze plates have
circular impressions filled with hard graphite (lubricant). Rockers are generally made of A-27 cast
steel.

Expansion Bearings

Expansion bearings are positioned to be centered or plumb at a neutral temperature. That neutral
temperature is generally 68 degrees F for most railroads, but may be at a lesser temperature for
northern regions or higher for hotter climates.

Sliding Plates

This bearing type is used for short spans where span deflection is not a concern. These simple
bearings generally rest on a thick masonry plate or tower top and are anchored at the fixed end by
anchor bolts through the sole plate in a hole of about 1/8 in. oversize. At the expansion end, the sole
plate is slotted with the anchor bolt within the hole to provide lateral resistance. For longer spans
sometimes the plate is beveled or a semi-curved surface sandwiched between plates to allow for
some modest span rotation. Under the expansion sole plate may be a bronze plate to reduce thermal
movement friction. For some applications, teflon sheets are used for low friction slip surfaces in
sliding bearings.

Common defects:
• sole plates deformed from lack of support.
• base plate pounded into pier top.
• thin plate worn from abrasion in sliding surface.
• anchor bolts deteriorated at pier surface.
• sole plate at end of travel in anchor bolt slot.
• bearing buried in debris.
• lead or cotton duck sheet expelled, teflon sheet torn or expelled.

Cylindrical rollers (nested)

Bridges, circa 1900, often had cylindrical roller nests located between a masonry plate and an upper
casting with that casting pinned or hinged to allow for span deflection/rotation. The rollers generally
were in the 3 in. to 6 in. diameter size and the parallel rollers of the nest were connected on their
ends by a common bar with a shoulder bolt tapped into the end of each roller, which unitized the
group. The rollers were virtually line contact loading between the upper casting and lower plate. A
number of those roller bearings remain in use today and will be encountered by the inspector.

Common defects:
• side bar corroded and separated from rollers.
• rollers skewed, out of position, or missing .
• roller imprint in base plate from overload and/or cold working.
• restricted roller movement from debris or corrosion.

Segmental rollers (nested) - as the name implies, the rollers are segments of rollers. To some degree
they are rollers with trimmed sides which allow for more rollers to fit closer together in a group.
Segmental rollers are the most common type roller used in conjunction with truss spans. The roller is
actually somewhat of rectangular shape in section with a larger radius than the depth of the roller to
allow for more surface contact top and bottom and for a larger spread of load to the base. Segmental
roller nests usually have a pair of side bars that keep the nest of rollers uniformly spaced and
functioning as a group. These bars are bolted to the ends of the rollers. One style of segmental rollers
uses pintles that are inserted into both the top casting and masonry plate and engage each roller in a
socket. This style roller is forced to function by virtue of being engaged by the pintles. The pintles
also act to transfer lateral forces between the top shoe and base.

The more common style of segmental roller used in the railroad industry has side bars, but in lieu of
having pintles, there is a keyway raised in the top casting and masonry plate and a depressed slot in
each roller for engaging the components. This keyway is orientated longitudinally and located along
the centerline of the rollers. Like the pintles, the keyway assists in transferring lateral forces from the
top casting to the base.

Common defects:
• segmental rollers locked in an expanded position.
• segmental rollers frozen from side bar corrosion.
• side bar bolts broken.
• base plate with imprint of rollers.
• rollers with flat contact surfaces.
• segmental rollers in improper position for given temperature.
• pins between bearing upper and lower components rusted and restricting component rotation.
• debris/corrosion under rollers, restricted movement.

Rockers

A rocker assembly consists of an upper casting attached to the span, a common pin connecting the
component to the rocker, and the rocker itself. A single rocker is used in the assembly which
generally flares outward from the pin to the base. The base has a moderate radius which allows for a
moderate contact footprint on the base. The rocker can usually handle significantly larger
longitudinal span movements than a roller, but normally does not have as much load capacity as a
series of rollers. Rockers generally have pintles between the rocker and the masonry plate with the
pintle inserted in the masonry plate and a recess in the rocker. Rockers are more commonly
associated with girder spans on masonry piers.

Common defects:
• rocker with incorrect inclination.
• dirt under rocker, restricting movement.
• pins between components rusted and restricts component rotation.

Pot Bearings

This type bearing is relatively new in the railroad industry. Its main feature is that it allows for
multi-directional movement.

Common defects:
• dust skirts deteriorate.

Rubber/Neoprene Bearings

These type expansion bearings are generally only used in newer short steel span applications.

Common defects:
• overload deforms the material and flattens and/or tears the material.

Fixed Bearings

Fixed Position Sole Plate

These bearings are generally used on short spans where span deflection/rotation is not a concern.
These bearings are commonly used for beam spans and girder spans atop towers. Anchor bolts
through minimal clearance sole plate holes hold the bearing in place.

Common defects:
• sole plate pounded into pier top.
• sole plate deformed due to lack of support.
• anchor bolts deteriorated.

Fixed Shoes

This generally consists of an upper and lower casting connected by a central pin which allows for
span deflection/rotation. The fixed bearing may be a castings or a weldment. The lower casting is
held in position by anchor bolts through the corners of the base.

Common defects:
• pin corroded and does not allow rotation.
• anchor bolt deterioration.
• lower casting tipped by uneven embedding into bridge seat.

Bolsters

These are the equivalent of concrete blocks placed under bearings as spacers. Bolsters may be
castings, built-up sections or weldments. Bolsters should be carefully orientated so that the bolster
stiffeners can sufficiently resist applied forces.

Common defects:
• bent or cracked stiffeners.
• lack of anchorage to support.
• base pounded into pier top.

BEARING INSPECTION CHECKLIST

Expansion Bearings

Free of debris, corrosion.


Evidence of unrestrictive movement.
Correctly positioned for given temperature.
Uniform, supported on base material.
Anchor bolts present, full section and tight.
Central bearing pin allows span deflection/rotation.
Pintles, keyways, side bars present and engaged.
Fasteners to superstructure full sections and tight.

Fixed Bearings

Free of debris, corrosion.


Uniform, supported on base material.
Anchor bolts present, full section and tight.
Central bearing pin allows span deflection/rotation.
Fasteners to superstructure full sections and tight.

Bolsters

Free of debris, corrosion.


Uniform, supported on base material.
Anchor bolts present, full section and tight.
Stiffeners free of bents and cracks.
Fasteners to bearing full sections and tight.

Emergency

If the inspector discovers a bridge condition that affects the integrity of the bridge under train loads,
contact the railroad dispatcher and responsible authority to stop any trains and arrange for immediate
repairs.

Criteria for the limits of acceptable/tolerable conditions must be provided by the Railroad Owner or
their designated engineer based on the principles of acceptable rating contained in the appropriate
chapter of the AREMA Manual for Railway Engineering.
Expansion Bearings

Figure 9-23 Sliding Plate Figure 9-24 Left side expansion (slotted hole)
Nomenclature Right side fixed (tight hole)

Figure 9-25 Circular Roller Figure 9-26 Circular Roller


Nomenclature

Figure 9-27 Segmental Roller Figure 9-28 Segmental Roller


Nomenclature Arrow – roller flat spots
Expansion Bearings

Figure 9-29 Rocker Figure 9-30 Rocker


Nomenclature Note excessive lean

Figure 9-31 Sliding plate Figure 9-32 Segmental roller


Bearing pounded into pier Loose granite block

Figure 9-33 Segmental Roller Figure 9-34 Saddle


Critical condition
Fixed Bearings

Figure 9-35 Sliding Plate Figure 9-36 Flat plate


Nomenclature

Figure 9-37 Fixed Bearing Figure 9-38 Fixed bearing


Nomenclature

Figure 9-39 Bearing buried in debris Figure 9-40 Loose and missing fasteners in internal
diaphragms
Fixed Bearings

Figure 9-41 Fixed bearing – flat plate Figure 9-42 Bolster


Depressed into pier Fractured Beam

Figure 9-43 Cracked casting Figure 9-44 Flat plate with allowance for span
deflection

Figure 9-45 Anchor bolts Figure 9-46 Bearing base not fully supported
Mispositioned due to pier movement
SECTION 4
STEEL BEAM SPANS

General

Timber, concrete, and steel beam spans are the most common bridge spans in the railroad network.
Within the network, some railroads prefer steel to timber for beam spans in the shorter lengths such
as that used for trestle applications. Others prefer steel for its long life, and in favorable climates for
the minimal maintenance needs and for ease of repair.

Beam spans are often used to span small streams, used in overpasses and underpasses because of
their shallow depths, and in trestles.

Rolled, welded or built-up section, steel beam spans are generally economical to 50 ft. in length. The
limiting factor is deflection of the spans under load which is limited to 1/640 of the span length.

Features

Steel beam spans are generally composed of two or more parallel beams per track with diaphragms
and bracing as required. The spans also generally have flat plate bearings with one end fixed and one
end having slotted holes for the flat plate to slide, accommodating expansion requirements.

Beam spans may carry either an open deck where the ties are directly in contact with the top flange
of the beams or ballasted decks. The ballasted deck may be of timber, concrete or steel plating.

Beam spans, particularly shorter length spans, are subjected to significant cyclic loading from each
set of coupled trucks which fully load the span, followed by total unloading between sets of trucks.
Thus, the inspection for fatigue damage is an important item.
The inspector should be mindful of the following facts concerning beam spans and conduct the
inspection accordingly:

• a two beam span is fracture critical.


• a four beam span is redundant and not fracture critical.
• a rolled or welded beam is not internally redundant.
• a built-up beam of a web plate and bolted or riveted angle flanges is internally redundant.
• a built-up beam of a web plate and angle flanges and bolted or riveted cover plates is
internally redundant.
• a built-up beam of a web plate and angle flanges and WELDED cover plates is NOT
internally redundant.

Commonly Found Defects

• section loss under/adjacent to open deck ties.


• notches and other damage in bottom flanges from accidents.
• bearings and bearing plates pounded into masonry pier tops (particularly on concrete,
limestone or sandstone seats).
• section loss at the bearing areas.
• knife-edged bearing stiffeners at the bottom flange, over the bearings. In advanced
conditions - possible bottom flange-to-web crack behind the stiffener and beam from
rocking action.
• skewed spans - loose diaphragm fasteners and/or cracks in web at diaphragms.

Inspection Checklist

Track

Alignment - horizontal and vertical on approach, on structure, surface, clearances.


Support - behind abutment (hanging ties) and on span.
Fastenings to deck.
movement (sliding)
hook bolts
anchors on and off structure
General conditions.
ties
rails
joints
inner guard rails (if used)
outer guard timbers
spacer straps
spacer blocks
Walkways/Platforms

Walkway surface.
trip hazards
toeboards
fasteners, support elements, planking/grating
Handrail.
posts
rails
cables
fastenings
Overall safety items.

Beams
Top flanges (compression).
section loss (under/between ties)
cover plates (attachment to flange)
missing or worn fasteners
cracked welds both longitudinally and plate butt welds
cracks - in flange angle fillets and/or moon-shaped cracks in the flange
buckling from rail load impact
cracks from out-of-plane bending at diaphragms

Web.
buckling at beam ends from lack of stiffness
significant section loss within the end between bearing stiffener and first
intermediate stiffener
shear cracks in end section between bearing stiffener and first intermediate
stiffener
cracks from out-of-plane at diaphragms

Bottom flanges (tension).


section loss from corrosion
damage from underside strikes (vehicles/vessels)
missing fasteners/cracked welds

Bracing.
stiffeners at end supports
diaphragms - loose connections

Bearings.
functional - warped, worn, anchored
fixed bearing - stationary
expansion bearing - room for movement in slotted hole

Other Items.
span behavior to passage of live load - steady, excessive vertical or lateral
movement, deflection under load
clearance markings for highway underpass
utilities, attachments to beams
spilled ballast at abutment affecting beam span
cleanlines
steel protected, excessive local corrosion even at non-critical locations

Emergency

If the inspector discovers a bridge condition that affects the integrity of the bridge under train loads,
contact the railroad dispatcher and responsible authority to stop any trains and arrange for immediate
repairs.

Criteria for the limits of acceptable/tolerable conditions must be provided by the Railroad Owner or
their designated engineer based on the principles of acceptable rating contained in the appropriate
chapter of the AREMA Manual for Railway Engineering.
Beam Span

Figure 9-47 Simple spans Figure 9-48 Underside – Four beam system
Ballasted deck (redundant)
Note bracing system and diaphragms

Figure 9-49 Beam web loss at support Figure 9-50 Interior diaphragms
Note debris behind deteriorated area

Figure 9-51 Fatigue cracks in beam web due Figure 9-52 Damaged beam
to out-of-plane bending
SECTION 5
FLOOR SYSTEMS

General

Steel floor systems are commonly composed of stringers and floorbeams within the girder and truss
spans. They may support an open deck with ties placed directly on the stringer tops or there may be
ballasted decks of timber, concrete or steel construction atop the stringer and floorbeams. A concrete
ballast trough may be either composite or noncomposite with the stringers or floorbeams. There are
a number of other types of floor systems such as: closely spaced floorbeams with diaphragm spacers
often used to maximize vertical clearance in girder spans over highways, steel trough system, and
composite concrete/steel beam system.

Types of Floor Systems

Conventional Stringer-Floorbeam System

Conventional stringer-floorbeam systems are composed of stringers framing into floorbeams and
connected to girders or trusses, and generally have a panel length (floorbeam-to-floorbeam spacing)
in the range from 10 ft. to 15 ft. for girders and 25 ft. to 30 ft. for trusses. Both the stringers and
floorbeams may be rolled, welded or built-up sections with end connections generally being a pair of
bolted or riveted angles. A variation of the above system is to have simple span stringers supported
atop the floorbeams. This is more commonly found on deck truss type spans.

Floorbeam-Diaphragm Systems

A floorbeam-diaphragm system is a system of closely spaced, transverse shallow depth floorbeams,


spaced by small diaphragms. This system is generally used in conjunction with thru-plate girder
(TPG) spans where the distance from base of rail to underside of the girder span must be kept at a
minimum primarily due to existing site conditions. Steel ballast deck plating, bolted, welded or
without connection covers the floorbeam system, or direct fixation fasteners atop the floorbeams
may carry the rail. The shallow floorbeams are governed by span (girder-to-girder) length deflection
criteria.

Trough System

A trough system provides the least depth between base of rail and underside of floor system. This
somewhat outdated system is composed of a series of transverse plates and angles used to form
narrow U-shaped troughs sufficiently wide enough to place ballast and a tie such that the base of
rails are only a matter of several inches above the upper horizontal plate connection between
troughs.

Composite Concrete/Steel Floor System

A composite concrete/steel floor system is also used to minimize the depth of floor system between
base of rail and underside of the floor. The composite floor system likewise is composed of simple
spans either framing into floorbeam webs or supported atop the floorbeam with shear studs linking
the reinforced steel concrete slab to the stringer which usually is a rolled or welded section. In
minimizing deck depth, such composites may have direct rail fixation consisting of inserts placed in
the concrete for the attachment of tie plates.

FLOOR SYSTEM INSPECTION CHECKLIST

When inspecting a floor system, the inspector should be mindful of the following facts concerning
floor systems and conduct the inspection accordingly:

• A two stringer per track span system either framing into the web of a floorbeam or supported
atop floorbeams, is fracture critical.
• Floorbeams of a conventional floor system are fracture critical.
• A four beam per track stringer span between floorbeams is redundant and not fracture critical.
• A stringer or a floorbeam built-up of a web plate and angle flanges is internally redundant.
• A stringer or a floorbeam built-up of a web plate and angle flanges with bolted or riveted cover
plates is internally redundant.
• A stringer or a floorbeam built-up of a web plate and angle flanges with WELDED cover plates
is NOT internally redundant.
• A floorbeam-diaphragm floor system of closely spaced floorbeams is not fracture critical and is
generally considered redundant.
• A trough floor system is not fracture critical and is redundant.
• A composite floor system follows the same designations as if it were not a composite system.

FLOOR SYSTEM INSPECTION CHECKLIST

Track
Alignment - horizontal and vertical clearance on approach, on structure, clearance.
Support.
stringer top flange adequate
fastening to deck – movement (sliding)
bolts
hook bolts or equivalent
anchors on and off the structure
General conditions.
ties
rail joints
inner guard rails (if used)
outer guard timbers
spacer straps
spacer blocks

Walkways/Platforms
Walkway surface.
holes
trip hazards
toeboards
Support - connections to stringers.

Handrail
Posts.
Rail cables.
Fastenings.
Overall safety.

Conventional Floor System

Stringers and Floorbeams - Simple Spans

Top flanges (compression).


section loss - under/between ties (major losses are significant as related to beam buckling
and secondarily as related to support of ties)
cover plates - attachment to flanges; missing/worn fasteners, cracked welds both
longitudinally and plate butt welds
stringer cracks - in flange angle fillets and/or moon-shaped cracks in the flange

Web.
buckling due to a lack of stiffener bracing
significant section loss within the end distance from support and in floorbeam web
over stringer connection
shear cracks in end distance
out-of-plane cracks - at copes at stringer or floorbeam ends, at stringer diaphragms, in
floorbeam over stringer connection

Bottom Flanges (tension).


section loss - mid-span losses; stringer losses may control capacity of span, stringer
loss where bottom laterals cross the flange; floorbeams under stringer connections
cover plates – attachment

Stringer Bracing.
top laterals - section loss in angles and connection plates; deteriorated, loose or missing
fasteners, cracked welds
diaphragms - section loss in angles and connections, deteriorated, loose or missing
fasteners

Connections.
stringer web-to-floorbeam web - cracks in connection angle, in the fillet at either end or
in line with the line of fasteners in the floorbeam; loose and/or broken fasteners at
top or bottom of connection
floorbeam-to-girder or truss - loose or missing fasteners
stringers atop floorbeams - section loss in stringer end web stiffeners at bottom flange,
sole plate, floorbeam top flange, floorbeam stiffeners under top flange; deformed
supports; missing anchorage

Stringer Bearings.
stringers atop floorbeams - see Connections
saddles - (used as relief points in long spans) stringer sole plate wear, anchorage, loss of
section in saddle side web plates alongside saddle casting, floorbeam web loss behind
saddle, loose/missing saddle fasteners

Floorbeam-Diaphragm System (often used in conjunction with ballast deck)

Floorbeams.
top flanges (compression)
section loss - reduced capacity, buckling
deck connected or tight to floorbeam top flanges

Web.
significant section loss within “d” distance
shear cracks in end “d” distance
cracks - at diaphragm connections
out-of-plane cracks in web copes at beam ends and at diaphragm connections

Bottom Flanges.
section loss - mid-span
accident damage

Diaphragms.
section loss
cracks in diaphragms

Connections.
floorbeam-to-floorbeam - connection angles - section loss, cracks in angle fillet or in line
with fasteners; loose, broke or missing fasteners

Floorbeam-to-girder.
same as floorbeam-to-floorbeam

Trough System

System acts as a series of connected - winged U shaped transverse beams with the top
horizontal plate being the compression flange, the bottom of the trough horizontal
plate the bottom flange and the side vertical plates as the webs.

Section loss - laterally, mid-span in the bottom flange; anywhere from contact by ballast,
in web end “d” distance due to lack of drainage.
Cracks - low profile system used for underpasses - accident damage.
Connections - loss of fastener heads in contact with ballast.

Composite Concrete/Steel Floor System

A composite system allows the concrete deck to be part of the flange of the system and
thus reduces the size of the steel support beams. Note that many concrete decks
are not composite with the floor system. See bridge plans to determine whether
the deck is or is not composite.
Concrete deck - see Concrete Bridge Inspection.
Stringers.
see Conventional Floor System for inspection items
loss of top flange section from leakage through cold joints
Floorbeams.
see Conventional Floor System for inspection items
Connections.
pumping between the deck and support may indicate failure of shear studs and loss
of composite action
see Conventional Floor System for inspection items

Commonly Found Defects

1. Stringer-to-floorbeam connections - loose/broken fasteners in connection angle on


the floorbeam side.
2. Stringer-to-floorbeam connections - connection angle cracks in angle fillet top or
bottom of angle, or in line with end fasteners on floorbeam side.
3. Stringer bottom flange section loss where span bottom lateral system crosses the
flanges.
4. Stringer top flange losses adjacent to and under ties.
5. Floorbeam web losses over stringer connections.
6. Floorbeam web losses between span connection and stringer connection from
corrosion due to refrigerator cars having brine water dripping.
7. Stringer bracing connection plates and fasteners with section loss and disconnected.
8. Floorbeam top flange losses on underside of top flange at stringer connections.
9. Stringers atop floorbeams - poor seat areas, worn stiffeners, cracks in stringer bottom
flange-to-web at support, deformed seats.
10. Stringer web cracks in the “d” distance from out-of-plane bending in copes.
11. Stringer welded top cover plates with broken welds.
12. Floorbeam-diaphragm floor system with skewed ends - loose fasteners, diaphragm
connection cracks in skew area.
13. Stringer top flange cracks due to tie deflection and/or hook bolt pull-up.

Emergency

If the inspector discovers a bridge condition that affects the integrity of the bridge under train loads,
contact the railroad dispatcher and responsible authority to stop any trains and arrange for immediate
repairs.

Criteria for the limits of acceptable/tolerable conditions must be provided by the Railroad Owner or
their designated engineer based on the principles of acceptable rating contained in the appropriate
chapter of the AREMA Manual for Railway Engineering.
Floor System
Conventional Stringers - Floorbeams

Figure 9-53 Stringer floorbeam Figure 9-54 Stringer floorbeam


Ballast deck Open deck
Framed into side Stringers over floorbeams

Figure 9-55 Stringer Figure 9-56 Stringer


Typical top flange section loss Cracked top flange angle due to flexing of angle

Figure 9-57 Stringer top flange wear Figure 9-58 Stringer web losses
Floor System
Conventional Stringers - Floorbeams

Figure 9-59 Stringer web Figure 9-60 Stringer web crack with repair place –
Cracked at cope bolted

Figure 9-61 Stringer web Figure 9-62 Stringer


Crack above connection angle Web crack
Brittle fracture

Figure 9-63 Stringer bottom flange crack at Figure 9-64 Stringer diaphragm – broken
Connection to floorbeam connection
Floor System
Conventional Stringers - Floorbeams

Figure 9-65 Stringer web Figure 9-66 Stringer bottom flange


Potential crack locations Potential crack locations

Figure 9-67 Stringer bottom flange Figure 9-68 Stringer over floorbeam
Section loss at bottom lateral connection Bottom flange loss and poor floorbeam
Top flange

Figure 9-69 Floorbeam – top flange Figure 9-70 Corrosion and section loss under old
Crack grease paint
Floor System
Conventional Stringers - Floorbeams

Figure 9-71 Stringer - floorbeam connection Figure 9-72 Stringer – floorbeam connection
Upper fillet crack Loose and broken rivets

Figure 9-73 Stringer over floorbeam Figure 9-74 Stringer


Renewed support Bracing system

Figure 9-75 Stringer bracing Figure 9-76 Stringer bracing


Section loss and fracture at connections Connection plate deterioration
Floor System
Conventional Stringers - Floorbeams

Figure 9-77 Web crack at rough cut Figure 9-78 Web section loss above
Cope Stringer connection

Figure 9-79 Floorbeam web Figure 9-80 Floorbeam bottom flange


Section loss above stringer with cracks Section loss

Figure 9-81 Floorbeam web Figure 9-82 End Floorbeam


Crack square cope Stringer bracket – section loss
Floor System
Trough System and Composite Concrete/Steel

Figure 9-83 Trough system – deck level Figure 9-84 Trough system – underside of TPG
Note ties in trough pockets

Figure 9-85 Trough system – underside Figure 9-86 Trough system – advanced
General view of components deterioration due to lack of proper drainage

Figure 9-87 Composite concrete – steel Figure 9-88 Composite concrete – steel
Concrete and steel DPG act as single unit Concrete and steel (BMS)
Connected by shear studs
Floor System
Floorbeam – Diaphragm

Figure 9-89 Floorbeam – Diaphragm System Figure 9-90 Floorbeam – Diaphragm System
Three TPG spans, ballasted Difficult to see details between floorbeams

Figure 9-91 Floorbeam – Diaphragm system Figure 9-92 Floorbeam – Diaphragm system
Open deck – easy to view details Watch for cracked diaphragm
In skewed areas

Figure 9-93 Floor System – Box girder Figure 9-94 Floor System – Box girder
Direct fixation – underside Direct fixation - deck
SECTION 6
THROUGH - PLATE GIRDER (TPG) SPANS

General

Through-plate girder spans are economical in the 50 ft. to 150 ft. long range and used with both open
and ballasted deck applications. In recent years, with high strength steels, TPG’s in the range of 200
ft. to 250 ft. have been built. TPG spans are commonly used for highway underpasses, stream
crossings and where a low profile from rail to bottom of span distance is required.

Nomenclature

The standard nomenclature used by the majority of the railroad industry both for identification of
components, as well and the numbering system, is given in the “Nomenclature” section of this
handbook. Although the shown TPG span has built-up components, a welded span would have
similar component names.

Features

A unique feature of the TPG span is no upper bracing for the compression flange of the girder. The
only bracing for the girder are knee braces which provide rigidity between the floorbeams and girder
web at each panel point and a bottom lateral system. Some TPG spans have external knee braces
which also align with the floorbeams. The TPG generally has mirror image web stiffeners spaced
less than the depth of the girder “d” distance. The first web stiffener is generally located at a distance
of about “d/2” from the end bearing stiffeners. Heavier end bearing stiffeners are located directly
above the bearing.

The TPG span may have one of the following type floor systems:

• conventional floorbeam-stringer system.


• floorbeam-diaphragm system.
• trough system.
• concrete-steel composite.

For shorter TPG spans, the bearings generally consist of sole plates in fixed or sliding applications
for fixed and expansion bearing applications, respectively. For longer spans where there is a modest
degree of girder rotation or deflection, a curved sole plate bearing, a pinned bearing, either a roller or
rocker may be used. In all cases, the bearing system should be anchored to the supporting element
(pier, tower or pile bent).

A two-girder, TPG span is fracture critical in that there are only two girders. Should one girder fail,
or be crippled, the span is in danger of failure. A TPG span of built-up girder tension flange sections
has internal redundancy, i.e., should a crack start in a flange angle, the crack will not jump to the
other angle or web. Should there be welded cover plates on the tension flange of the same span, the
internal redundancy is no longer there. A welded TPG span is fracture critical and non-redundant.
Some through girders have been fabricated with welded compression flanges and bolted tension
flanges to provide internal redundancy.

The determination of fracture critical members and redundancy of the floor system of a TPG span is
discussed in the Steel Floor System section of this handbook.

TPG Inspection

The inspector should observe the TPG span in a global context:

• the whole span as a unit looking for span alignment with other spans.
• signs of settlement.
• signs of tilt or twist.
• span crowding between backwalls.
• bunching of multiple spans together.
• alignment of the track within the span.

Sometimes curved track is realigned and does not remain in proper position for horizontal clearance
on both sides. The bunching of spans is sometimes noted on high trestles and can be of concern
should there be a lack of adequate support for the span bearing atop a tower.

Inspectors should observe the whole span under live load conditions, noting excessive deflection,
sway or any other abnormal movement.

Skewed TPG spans are not desirable, but many are built to accommodate highway alignment and
stream alignment to the track. The skew end panel(s) has long and short members in parallel and
under the heavy rail loads, deflections are different and thus susceptible to connection fastener
failures and crack development in floorbeams and girder webs from out-of-plane bending,
particularly near diaphragms or brace frames between stringers.

The inspection of the TPG span should be in a methodical manner. If done by one person, the person
should establish a pattern such as left side from the near end to far end and return on the right side,
or do left then right side at each panel then proceed to the next panel. A pair of inspectors can
readily work in tandem, one inspecting the left side while the other inspects the right side.

In an open deck bridge, the inspector can step down onto the bottom lateral connection plates and
obtain a good view of the stringer-floorbeam connection as well as observe any stringer flange
section loss. A major portion of the bottom lateral system can be observed at the same time. In
ballasted deck bridges, generally the entire floor system is covered and must be viewed from the
underside from ladders or using an inspection vehicle.

The inspector should observe the knee braces closely for full connection to the girder web as well as
to the floorbeam top flange.

On ballast deck TPG’s where the ballast contacts the girder web, corrosion and associated section
loss may be found in the web. The inspector should routinely inspect his area by removing random
selected areas of ballast.

TPG INSPECTION CHECKLIST

Track
Alignment - horizontal and vertical on approach, on structure, surface, clearances.

Support.
behind abutment (hanging ties)
at end stringer brackets
stringer flanges

Fastening to deck.
movement (sliding)
hook bolts
rail anchors on and off structure

General conditions.
ties
rails
joints
inner guard rails (if used)
outer guard timbers
spacer straps
spacer blocks between ties

Walkways/Platforms

Walkways on TPG spans are commonly between the girder and track, from panel point-to-panel
point. Sometimes walkways are positioned to the outside of the TPG span and attached to the girder
stiffeners, particularly where there is a tight clearance. Platforms are usually at the ends of the span.

Walkway surface.
holes
trip hazards particularly at knee braces

Handrail.
posts
rails
cables
fastenings of walkway leading from approach to TPG span

Overall safety items.

Top flanges (compression)


section loss.
accident damage.
cover plates.
attachment to flange
missing/loose fasteners
weld at flange size transitions (change in thickness and/or width)
cracks - in flange-to-web
flange buckling from lack of knee braces

Web
section loss from corrosion where ballast contacts with web.
buckling from lack of stiffeners.
buckling from loss or damaged knee braces.
shear cracks in end “d” distance.
loose fasteners in web splices.
web cracks at welded connections.
web cracks at ends of partial depth stiffeners.

Bottom Flange
section loss in center portion of span.
stress concentrations from accident damage (nicks, scoring, bending).
cover plates.
attachment to flange
missing/loose fasteners
sheared fasteners from accident damage
weld at flange transitions (change in thickness and/or width)
cracks - flange-to-web, attachments, in bottom flange at bearing outline area

Knee Braces
section loss - primarily adjacent to connection to floorbeam .
cracks - connection angles to web in angle fillet.
fasteners - loose/missing rivets or bolts. Upper rivets often have deformed heads due to
the limited area for installing the rivet.
damage - disconnected from floorbeam due to shifted rail loads.

Bottom Laterals
section loss - at stringer and end connections.
spreading and section loss in vertical legs of those laterals with back-to-back angle
sections with excessive fastener spacing.
cracks - adjacent to end connections.
fasteners.
loose/missing fasteners at connection to stringer bottom flanges
deteriorated heads on end connection plates
missing sections from accident damage

Floor System
See Floor Systems section of this Chapter for additional inspection information.

Bearings
Sliding plate, rollers, rockers: See Bearings for additional inspection information.

Common TPG Span Defects

1. Damaged knee braces from wide loads.


2. Damaged bottom girder flanges and bottom laterals when spanning over highways.
3. Corrosion pockets with section loss to bottom flanges adjacent to bottom lateral
connection plates.
4. Girder end deterioration above the bearing area.
5. Section loss in bottom laterals and stringer bottom flanges where connected.
6. Broken and loose fasteners at bottom lateral-to-stringer connection.

Emergency

If the inspector discovers a bridge condition that affects the integrity of the bridge under train loads,
contact the railroad dispatcher and responsible authority to stop any trains and arrange for immediate
repairs.

Criteria for the limits of acceptable/tolerable conditions must be provided by the Railroad Owner or
their designated engineer based on the principles of acceptable rating contained in the appropriate
chapter of the AREMA Manual for Railway Engineering.
THROUGH PLATE GIRDER (TPG) SPANS

Figure 9-95 TPG – side view Figure 9-96 TPG – open deck – end view

Figure 9-97 TPG – common location for loose Figure 9-98 TPG – ballasted deck – end view
fasteners

Figure 9-99 TPG – damaged knee brace Figure 9-100 Deteriorated brace
Knee brace provides girder stability
THROUGH PLATE GIRDER (TPG) SPANS

Figure 9-101 Potential crack locations Figure 9-102 Potential crack locations
Short bearing and interior stiffeners From weld intersection

Figure 9-103 Potential crack locations welded Figure 9-104 Potential crack locations – welded
attachments connection plate

Figure 9-105 Bottom flange/web crack adjacent to Figure 9-106 Potential crack location – welded
welded stiffeners attachment to web
SECTION 7
DECK PLATE GIRDER (DPG) SPANS

General

Deck Plate Girder spans are economical in the 50 ft. to 150 ft. long range and used in both open and
ballasted deck applications. With the advent of high strength steels, DPG’s in the range of 200 ft. to
250 ft. have been built, primarily of welded construction. DPG spans are commonly used in
applications where the depth of structure from rail to bottom of span is not a factor. In fact, DPG
spans help reduce the overall height of piers and viaduct bents/towers by having a deeper section.
Additionally, DPG spans are ideal in that there are no clearance restrictions associated with the type
span.

Nomenclature

The standard nomenclature used by the majority of the railroad industry both for identification of
components as well as the numbering system is given in the “Nomenclature” section of this
handbook. Although the shown span represents a span of built-up components; a welded span would
have similar components and similar component identification.

Features

The DPG span generally does not have a floor system. For open deck the ties bear directly on the
top flanges. The ballast deck also rests directly on the top flanges. When the DPG span has multiple
cover plates, the open deck tie depths must be varied to accommodate the differences in heights.
Often the girders of a DPG span are spaced either at the 7 ft. to 9 ft. range. The wider the girder
spacing, the deeper and stronger the ties or deck must be to carry the rail load. The DPG, unlike the
TPG, has a top and bottom lateral bracing system and cross frames with top and bottom struts at the
span ends and intermediate points.

For shorter DPG spans the bearings generally consist of sole plates in fixed or sliding applications
for fixed and expansion bearing applications, respectively. For longer spans where there is a modest
degree of girder rotation or deflection, curved sole plate bearings, a pinned type bearing, of either a
roller or rocker may be used. In all cases, the bearing system should be anchored to the supporting
elements (abutment, pier, tower top or pile bent top).

A single-track, DPG span is fracture critical in that there are only two girders. Should one girder fail
or be crippled, the span would be in danger of failure. A DPG span of built-up girder sections has
internal redundancy, i.e., should a crack develop in one element, the crack will not migrate to the
other elements, except should there be a retrofit that welds cover plates to the girder flanges. A
welded DPG span is fracture critical and non-redundant.

Inspection

The inspector should view the DPG span from a global context before proceeding with the
individual elements of the span. Span alignment with other spans, signs of settlement, signs of tilt,
span crowding between backwalls, span bunching, and even laterally displaced DPG indications
should be investigated. The bunching of spans on long viaducts can be of major concern should a
span lack adequate support for the span bearing atop a tower.

Inspectors should observe the behavior of the DPG span under live load conditions, noting excessive
deflection, sway or any other abnormal conditions.

Skewed DPG spans are not desirable but often unavoidable due to the alignment of what is being
spanned. Skewed ends result in differences in deflections among elements and result in the breakage
of fasteners, cracking of connections and out-of-plane bending which results in fractures.

The inspection of a DPG span is rather simple - observing the outside surfaces of the girders and
walking through the interior of the span either by walking the bottom laterals or from a walkway
supported on the lateral system. Some bridges have a hand cable through the girder stiffeners for the
inspector to walk the girder bottom flange, or a safety cable through the cross frames for the
inspector to attach a harness lanyard.

DPG INSPECTION CHECKLIST

Track
Alignment
horizontal on approach
vertical on approach
on structure
surface

Support
behind abutment (hanging ties)
on girder top flanges

Fastening to girder flanges


movement (sliding)
hook bolts
rail anchors on and off structure

General condition
ties
rails
joints
inner guard rails (if used)
outer guard timbers
spacer blocks between ties

Walkways/Platforms

Walkways commonly are between the rails (in the gage) or between the tracks or bracketed from
extended ties. Platforms are used as refuge areas at a given spacing along a bridge.

Walkway surface
holes
tripping hazards
toe boards
fasteners, supports, deck material (plumbing/grating)

Handrail
posts
rails
cables
fastenings

Overall safety items

Top Flanges (compression)

section loss
under/between ties, major losses are significant as related to girder capacity, buckling
and secondarily as related to tie support
cover plates
attachment to flange
missing fasteners
worn fasteners
cracked welds, longitudinally and at plate butt welds, transitional flange
thickness/width welds
cracks - in flange angle fillets and/or moon-shaped cracks in the flange from tie
deflection to the gage side of the girder and/or pulling up on the field side from the
hook bolts raising as the ties deflect. Generally the cracks occur when the flange is
only a pair of angles and less than 3/4 in. thick.

Web
buckling at girder ends from lack of stiffness
significant section loss within the end “d” distance
shear cracks in end “d” distance
out-of-plane cracks in web at girder ends from lateral movement, at cross-frames, at partial
depth stiffeners
splices – fasteners

Bottom Flange (tension)


section loss - mid-span losses generally control capacity of the span
cover plates - weld attachments, damage, transitional flange thickness and width welds
cracks - at attachments, accident damage

Bracing
web bearing stiffener loss of support at bottom flange
web intermediate stiffeners
cross frames and struts - section loss to angles, particularly back-to-back angles with
excessive fastener spacing, cracks in connections to girders, connection plate losses
bottom and top laterals - section loss in angles adjacent to connection plates and swelling
between back-to-back angles, loss of fasteners in horizontal connection plates,
corrosion losses in horizontal connection plates

Bearings
functional - accommodates expansion and girder rotation
fixed bearings are stationary
expansion bearings in correct position at a neutral position with room for movement

Other Items
span behavior to passage of live loads - steady, excessive vertical or lateral movement,
deflection under load
clearance markings for highway underpass
utilities, attachments to girders
cleanliness
steel protection, excessive local corrosion even in non-critical areas

Commonly Found Defects

1. Section loss in girder top flanges under/adjacent to ties.


2. Fractures, moon-shaped in girder top flange from out-of-plane bending due to tie deflection.
3. Section loss in girder bottom flange and web at bearing areas.
4. Knife-edged bearing stiffeners at the girder bottom flange, over the bearing. In advanced
conditions possible bottom flange-to-web crack behind the stiffener due to beam
rocking.
5. Skewed spans - loose diaphragm fasteners and/or cracks in cross frames.
6. Deteriorated lateral bracing members adjacent to girder connection plates.
7. Bracing strut section loss adjacent to girder connection.
8. Girder bottom flange losses adjacent to bottom lateral connection plates.

Emergency

If the inspector discovers a bridge condition that affects the integrity of the bridge under train loads,
contact the railroad dispatcher and responsible authority to stop any trains and arrange for immediate
repairs.

Criteria for the limits of acceptable/tolerable conditions must be provided by the Railroad Owner or
their designated engineer based on the principles of acceptable rating contained in the appropriate
chapter of the AREMA Manual for Railway Engineering.
Deck Plate Girder (DPG) Spans

Figure 9-107 Plate girder on tower Figure 9-108 DPG – side view

Figure 9-109 DPG – double girder underside view Figure 9-110 DPG – loose cross-frame connection
of bracing system

Figure 9-111 DPG – top flange crack from tie Figure 9-112 DPG – ends of girders atop tower –
deflection (out of plane bending) debris hides details
SECTION 8
TRESTLES AND VIADUCTS

General

Steel trestles and steel viaducts make up a significant portion of the railroad industry’s bridges.
Trestles generally refer to structures composed of beam spans supported on steel pile bents. These
structures commonly support elevated track found in urban areas, where embankments are not
practical or in rural areas where there may be a wide flood plain requiring significant open space
beneath the track. Viaducts generally refer to structures composed of beam, girder and deck truss
spans or a combination thereof, supported on steel towers, such as that used to cross a ravine.

The distinction between trestles and viaducts is related to the length and arrangement of spans, not to
height. The generally greater height of viaducts vs. trestles is a result of using the type of structure
that provides the required height at minimum cost.

The term “trestle” has, in the North American railroad industry, been applied to structures with
relatively short (10 to 35 ft.) spans of essentially uniform length supported on driven or framed
bents, regardless of height, which sometimes exceeds 100 ft. Trestles have bracing between bents
appropriate to the height. The term has been applied to timber, concrete and steel structures.
Trestles are usually constructed according to the railroad’s standard plans.

The term “viaduct” has been applied, particularly in Europe, to bridges of various materials and
types of construction whose primary function is to cross a deep, relatively wide valley at a
considerable height rather than to cross a major waterway. However, the term “viaduct” as applied
to steel railroad bridges in North America is defined in the 1920 AREA General Specifications for
Steel Railway Bridges as usually consisting of alternate tower spans and free spans of plate girders
or riveted trusses supported on bents with the tower spans usually being not less than 30 ft. long. In
practice, the spans between towers are usually at least twice the length of the tower spans.

Discussions of Beam, Through-Plate Girder, Deck Plate Girder and Deck Truss spans which are
supported by bents or towers are given in other sections of this Handbook, and will not be repeated
herein.

Nomenclature

The standard nomenclature used by the majority of the railroad industry for steel trestles and towers
is given in the “Nomenclature” section of this Handbook. Trestle bent and viaduct tower numbers
usually increase in the direction of ahead (increasing) mile post. Piles within a pile bent are
numbered in a left-to-right or reverse order, depending on a particular railroad’s established
numbering sequence, while facing the ahead direction. Tower legs are designated as left or right for
a given tower number. Tower tiers or stories are usually numbered from the top downward.
Features

Common Types of Steel Trestle Bents

• A transverse row of steel piles driven to a supporting elevation which are capped and braced
to support a span atop the cap. This type bent may have the outer piles battered for resisting
transverse loads. At a given spacing there may either be a pile bent consisting of a double
row of piles with transverse and longitudinal batters or a single row of piles with alternating
longitudinally battered piles. The latter two type pile bents are required to resist longitudinal
forces such as traction, braking and wind.

• A fabricated transverse frame of two columns with a base strut, a webbed cap beam and
diagonal cross bracing, atop concrete pedestals. At a given spacing along the trestle, there
will be a need for a bent which will resist longitudinal forces as described above.

Steel towers each resist a portion of the longitudinal and transverse imposed loads. The use of
DPG’s and/or deck trusses both help in reducing the required height of the tower which results in a
construction cost savings. Steel towers may be of several tiers in height. The common tower is
supported by either four independent concrete pedestals or by two piers. The longitudinal sides are
usually on a batter to give stability with the transverse sides being plumb. For long viaduct
approaches on a grade such as an approach to a waterway crossing, the tower tiers often are
fabricated in repeated sections, with the pedestal heights and one tier as the variables for grade
adjustments. A tower will typically support a DPG of approximately 40 ft. to 45 ft. in length. The
base of the tower may be a rigid, fixed section or it may be a hinged, pinned arrangement. Tower
tops are either individual flat plate platforms for the girder bearings or a header to support the
bearings. A header may be a single or double webbed girder orientated transversely to the track with
a top flange cover and a bottom flange of laced angles with batten plates. Towers are braced on all
four sides with either a V-shape or X-shape bracing pattern. There may or may not be horizontal
struts on all four sides at a common level in conjunction with the bracing system.

TRESTLE AND TOWER INSPECTION CHECKLIST

Accident Damage

The piling of a bent and column legs of a tower are compression members. As such, buckling is the
key item of concern. Accident damage to piles or tower members that would promote buckling
and/or the reduction of effective bracing within either a pile bent or a tower, both individual tower
members or the tower itself are items of serious concern.

Bents

Piling and Pile Bents

Section loss
affect pile capacity
may cause section buckling
loss at base affects stability
loss of longitudinal and transverse load resisting capacity

Columns (2 legs)
loss of section - affect capacity, section buckling
accident damage

Cap
unitizes the group of piles
provide adequate support for the span bearings

Bracing
section loss at connections
loose/missing fasteners

Loading
bent response to live loads

Towers

Columns
section loss - affects capacity, section buckling
accident damage that might promote buckling

Cap
Girder
functions as a beam between tower columns
section loss in top flange, too thin for bearings
top flange adequate for bearing support

Individual
sufficient section for support of span bearings
stiffeners support the cap plate
deformed cap plate

Base
Section loss - affects capacity and buckling
Anchorage - battered columns are anchored

Bracing
Bracing members in full section
Bracing solidly connected to connection plates

Loading
Tower response to live loads

Common Bent and Tower Defects

Bents
1. Section loss at pile base at ground line
2. Column leg corrosion and holes in base section
3. Bearing area on cap deformed
4. Bracing section loss - at connections and in members (lacing bars)

Towers
1. Section loss at column base
2. Lack of base anchorage
3. Deformed caps under bearing areas
4. Lack of adequate bracing supporting cap bearing area
5. Cap plate too thin
6. Cap connection cracks from girder span rigid connection to cap
7. Section loss from former fire barrel or brine leakage onto column legs and bracing
8. Loss of bracing lacing bars

Emergency

If the inspector discovers a bridge condition that affects the integrity of the bridge under train loads,
contact the railroad dispatcher and responsible authority to stop any trains and arrange for immediate
repairs.

Criteria for the limits of acceptable/tolerable conditions must be provided by the Railroad Owner or
their designated engineer based on the principles of acceptable rating contained in the appropriate
chapter of the AREMA Manual for Railway Engineering.
Trestles and Viaducts
Towers

Figure 9-113 Viaduct – general view Figure 9-114 Towers – General view
Note towers with inclined legs Note straight legs of double track bridge

Figure 9-115 Tower – circles indicate primary area Figure 9-116 Tower top – exterior side DFP riveted
for inspection tower cap

Figure 9-117 Rehabilitated tower cap – new cap Figure 9-118 Tower cap – note cap plate deflected
plate
Trestles and Viaducts
Towers and Bents

Figure 9-119 Tower cap – arrows point to crack in Figure 9-120 Tower base – note deteriorated
cap connection anchor bolt, left

Figure 9-121 Tower base Figure 9-122 Tower – damaged brace


Note holes in both sides above interior fill

Figure 9-123 Pile bent – check for section loss Figure 9-124 Pile bent – check pile-to-cap welds
above concrete jackets
SECTION 9
TRUSSES

General

There are many types of truss bridges. Only the most common type truss spans used throughout the
railroad industry are discussed herein. Unique trusses generally will require special instructions for
the inspector. The floor system (stringers and floorbeams) for trusses is discussed in the Floor
System section.

Through Truss

Through trusses where the track is located within the truss. The through truss span, like through
girder span, is predominantly used when it is necessary to minimize the underclearance of the area
being spanned such as a highway or stream.

Deck Truss

Like the deck girder span, the deck truss is used when the clearance beneath the truss is not a
governing factor. The deck truss is preferred where it can be used since it has unrestricted horizontal
and vertical clearance at track level. In addition, the depth of the truss reduces the overall height of
the piers or towers (economics).

There are many truss designs. Some truss types were patented and many bear the name of the
originator of that style. Some of the more common types of trusses are discussed and shown herein.

Nomenclature

The standard nomenclature for truss spans as adopted by AREMA is given in the “Nomenclature”
section of this Handbook. Specific railroads may use varying number systems, such as stringers
numbered from left-to-right or right-to-left, and tower tiers from top-to-bottom or bottom-to-top, and
so forth. Many inspectors have slang names for a bridge number elements which sometimes makes
it difficult in communicating a field problem to the office. Inspectors should be encouraged to use
standard nomenclature.
Truss Types

Warren Truss

Named after James Warren, the Warren truss with verticals is probably most commonly used in the
railroad industry. Warren trusses range in length from 100 ft. to 750 ft. with the more common
lengths being in the 150 ft. to 300 ft. range. It presently remains the basic design style. In the
Warren truss, the diagonals alternate in orientation within the truss. Generally in any one joint the
diagonal members will be alternating in tension and compression. The web system of the truss forms
a series of triangles. Most Warren trusses have verticals between the diagonals to limit the length of
the floor system panels and the unsupported length of the top chord. The verticals alternate in being
tension members (hangers) and compression members (posts) which carry insignificant load in a
through truss but full live load in a deck truss. The primary fatigue sensitive members in the through
Warren truss are the hangers which receive full tension live load followed by no live load (stress
cycle) for every set of wheels passing the panel point.

Warren

Pratt Truss

Named by Thomas and Caleb Pratt, the Pratt truss was often used in the 1900's to 1920's in the 150
ft. to 250 ft. length range. The unique feature of these trusses is that all interior diagonals are tension
members and are eyebars. Thus the truss weight is light and all truss joints are pinned, making
erection of the spans simple. The Pratt truss has many fracture critical members, and many have
significant maintenance problems. Because of the uniqueness of the Pratt pin-connected truss and its
associated problems, an entire Section is presented on Pratt Trusses.

Camelback

Parker

Pratt

Lattice Truss

This truss is a double intersection truss. It has parallel top and bottom chords with two or more sets
of diagonal (web system) members without vertical members, and without connection to each other.
These spans were built prior to 1900 and some were constructed of iron. A number of these remain
in main line service.

Howe Truss
Double intersection or lattice

Baltimore Truss

A truss generally used for large bridge spans up to 600 ft. long. This truss type has parallel chords
and sub-diagonals and sub-verticals which are used to keep the floor system panel lengths at a
reasonable length.
Baltimore

Pennsylvania Truss

A truss generally used for large bridge spans up to 600 ft. long. The truss style has a top chord and is
in cord segments making an arc, similar to a camelback truss, but with the chord sections generally
having the same cross-sectional make-up throughout its length. The truss has sub-diagonals and sub-
verticals which are used to keep the floor system panel lengths at a reasonable length.

Pennsylvania

Cantilever Truss

The cantilever truss bridge may be either a through truss or deck truss. Its application is generally
used to span great distances through the use of a truss span suspended from a cantilevered portion of
the bridge which is balanced by an anchor span. There are several variations, such as the use of a
partially suspended span used in lieu of the anchor span. This variation allows for the repetitive use
of identical spans. Spans’ lengths between main piers have reached 800 ft. Cantilever bridges have
been used to span major waterways and major ravines.

Cantilever bridges are also the choice where significant span lengths are needed plus the soil
conditions are poor. A cantilever bridge articulates at the hanger points without causing any distress
in the metalwork from pier settlement.

Cantilever
Pony Truss

A pony truss is analogous to the Through Plate Girder (TPG) span, but it has no upper bracing
system. The pony truss may be of several type trusses such as a Pratt or Warren. The truss top chord
at each panel point is braced with knee-type braces, either to the interior or exterior side of the
vertical.

Arches

There are basically three type arch spans:

• the through arch with the crown of the arc above the track with the floor system hung
from the arch in tension hangers.
• the deck arch with the floor system atop the arch with bents (spandrel bents) between the
deck and arch.
• the tied arch where the arch ends at the bottom chord and the bottom chord acts as a
tension member to resist the thrust of each arch. The merits of fixed or hinged arches are
not discussed, only the inspection of a hinge is presented.

Arch
Features

The bridge span referred to as a truss span, consists of the main carrying element a left and right
truss composed of top and bottom chords, vertical and diagonal members. The two trusses are
separated by:
• a floor system which can be at either the top chord or the bottom chord, and
• a bracing system

A simple span truss is similar to a beam: the top of the truss (top chord) is in compression like the
beam top flange; and the bottom of the truss (bottom chord) is in tension like the beam bottom
flange. The truss verticals and diagonal members are often referred to as the web members, similar
to the beam web. The train live load is carried by the floor system members (stringers and
floorbeams) to truss panel points where the loads enter the truss. Truss members carry live loads
axially along the length of each member to truss joints where the load is transferred to other
members. The ends of truss members may be pinned or fixed.

Early engineering theory readily understood axial forces, and resolved these forces at the truss panel
points by using pin connections. Additionally, pins simplified field erection; all members of a joint
were aligned and with the insertion of a pin, the connection was complete. Fixed truss connections,
commonly referred to as gusset connections, may require further analysis for load distribution
through the connection. Although the axes of the members intersect at a common point, the rigidity
of the joint introduces secondary stresses due to bending in the members as the truss deflects under
load. A gusset connection generally is connected to one of the joint members at the time of
fabrication and then other members of a joint are field riveted or bolted upon proper alignment. Field
welded connections are not permitted.

Trusses have fracture critical members (FCM). These members are tension members whose failure
would probably cause a portion of, or the entire bridge to collapse. These members need to be
identified for, or by the inspector. Those members should be closely inspected (within hand-range)
looking for defects or conditions that could cause the initiation of a crack. FCM with notches from
corrosion, tack welds left from fabrication and score marks are possible locations for the initiation of
a crack. Welded members do not have internal redundancy and thus are more susceptible to failure
from the initiation of a crack.

The cyclic loading of members is acute for truss hanger members, with each railcar imposing a stress
cycle on such a member. A single, heavily loaded coal train may impart 100 or more stress cycles on
a hanger while only imparting one stress cycle on a bottom chord. Thus, hangers are the most fatigue
sensitive member of truss members. In the floor system, the stringers and floorbeams likewise are
fatigue sensitive members - see the Floor System section of this Handbook.

The truss span must be able to withstand many imposed loads other than just the dead load of the
span, live load and impact from the live load. Those loads act in multiple directions on the span.
Some of these loads include wind, wind on live load, centrifugal load, nosing, earthquake and so
forth. A bracing system is used to resist, or at least participate in transferring these loads to other
bridge elements. The bracing system is often referred to as secondary members.

The key bracing systems are the top and bottom lateral systems which act as horizontal trusses to
carry the wind and other lateral loads imposed on the truss members to the ends of the spans where it
is transferred into the substructure. Since the wind can be in any direction, the lateral systems
commonly are arranged in an X-pattern with either member of the X being able to carry the load in
tension. For most trusses the bottom lateral system is connected to the stringers to reduce the overall
unbraced length of the member and to prevent it from buckling as it carries compression loads. Since
the wind load and other transverse, horizontal loads must be accommodated by the top and bottom
lateral system and that load transferred to the ends of the span, it will be noted that the members
usually increase in size from the center of the span to the ends. The upper lateral wind loads at the
end of the span must be transferred to the span base through the end posts; one of the reasons that
the end posts usually are larger or heavier members than the other truss members.

Other key bracing members of a truss are the end portal and the interior sway frames which give
rigidity to the span. The end portals are the primary elements for whole span rigidity plus they must
brace the end post. Because of the necessary size of end portals, many have been modified to
accommodate taller car loads. Most sway frames consist of an upper strut member directly
connected between the two truss top chords at a panel point along with a corner diagonal bracing
arrangement between the strut and a vertical member, usually a post.

TRUSS INSPECTION CHECKLIST

Track

Alignment
horizontal on approach
vertical on approach
on structure
surface

Support
behind abutment (hanging ties)
on span

Fastening to deck
movement (sliding)
hook bolts
rail anchors on and off structure

General condition
Ties
Rails
Joints
inner guard rails (if used)
outer guard timbers
spacer straps
spacer blocks

Walkways/Platforms

Walkway surface
holes
tripping hazards
toeboards
fasteners, support members, decking (planks/grating)

Handrail
posts
rails
cables
fastenings

Overall safety items

Truss Members

Top Chords
section loss at critical locations
adequate bracing for top chord and chord member internally
cracks - at chord splices, primarily at hip connection
wear in web pin holes
fasteners - loose, head loss

Bottom Chords
section loss at critical locations
eyebars equally tight, spreading from pack rust, section loss in head or body
pin - wear, scored, corrosion under spacer/collars
cracks
splices - fastenings, corrosion

Hangers
section loss in body, above connection to floorbeam
cracks - horizontally orientated at upper truss connection and commonly found at
lower row of fasteners
stress concentrations - tack welds, welded attachments, corrosion on edges
accident damage

Posts
alignment – buckling, accident damage
bracing – internal
member ends - pin hole wear, fasteners

End Post
alignment - buckling, accident damage
bracing - internal and external
section loss

Diagonal
section loss
compression member - internal bracing, alignment, accident damage
end connections - fasteners, pin hole wear
tension member - tight (equal tension in eyebars)

Bracing Systems

Top Laterals
section loss
connections
rod system – threads, turnbuckles, end loops/pins

Bottom Laterals
section loss - at end connections, at stringer connections, swelling within members
connections - at truss, at stringers
rod system – threads, turnbuckles, end loops/pins

Sway Frames
section loss
connections - to top chord and verticals
fatigue cracks in connections
rod system – threads, turnbuckles, end loops/pins

End Portal
section loss
accident damage
connections

Floor System - see Floor System section of this Chapter for inspection details.

Bearings - see Bearings section of this Chapter.

Commonly Found Truss Defects

1. Loose or unequally tensioned eyebars - diagonals, counters, bottom chords.


2. Worn pin holes in members.
3. Accident damage.
4. Deteriorated, loose, broken fasteners.
5. Cracks in hip joint connections.
6. Section loss from corrosion.

Emergency

If the inspector discovers a bridge condition that affects the integrity of the bridge under train loads,
contact the railroad dispatcher and responsible authority to stop any trains and arrange for immediate
repairs.

Criteria for the limits of acceptable/tolerable conditions must be provided by the Railroad Owner or
their designated engineer based on the principles of acceptable rating contained in the appropriate
chapter of the AREMA Manual for Railway Engineering.
Trusses

Figure 9-125 Trusses – deck cantilever Figure 9-126 Trusses - Pennsylvania

Figure 9-127 Trusses – arch Figure 9-128 Trusses - Baltimore

Figure 9-129 Trusses – Parker Figure 9-130 Trusses – Camelback


A form of Pratt truss A form of Pratt truss
Truss

Figure 9-131 Trusses – Warren –side Figure 9-132 Trusses – Warren – end view

Figure 9-133 Trusses – Warren – deck truss Figure 9-134 Trusses – double intersecting
Center span

Figure 9-135 Trusses – Pony – Warren Figure 9-136 Trusses – Pony - Pratt
Truss

Figure 9-137 Truss – check for changes in side Figure 9-138 Trusses – end post
clearance Commonly damaged by shifted loads

Figure 9-139 Trusses – truss hanger Figure 9-140 Trusses – hanger


Fatigue sensitive and fracture critical Section loss at critical section above floorbeam

Figure 9-141 Trusses – hanger crack in lower row Figure 9-142 Trusses – hanger with square cut
of fasteners of the U1 joint cope
Arrow points to crack
Truss

Figure 9-143 Trusses – top lateral system and Figure 9-144 Trusses – bottom lateral system
sway frame (background)

Figure 9-145 Trusses – top lateral system (rods) Figure 9-146 Trusses – top chord and lateral
system. Note excellent conditions without paint.

Figure 9-147 Trusses – Sway frame with build up Figure 9-148 Trusses – bottom lateral connection to
strut members and rod braces. stringer commonly found loose or broken fasteners.
Truss

Figure 9-149 Truss – bottom chord with section Figure 9-150 Trusses – bottom chord eyebars
loss damaged from debris during high water

Figure 9-151 Trusses – eyebar coming off pin Figure 9-152 Trusses – pin corrosion and collar
deterioration

Figure 9-153 Truss swelling from pack rust Figure 9-154 Trusses connection plate corner lifting
between poorly spaced fasteners from pack rust. Poor corner detached
SECTION 10
PIN-CONNECTED PRATT TRUSSES

General

This special section is provided because pin-connected Pratt trusses are unique and have many
fracture critical members with significant maintenance problems.

The Pratt truss is easily recognized by the diagonals, typically eyebars, sloping downward toward
the center of the span. The most common Pratt trusses are through trusses of six to eight panels in
length of 25 ft. long for an overall length from 150 ft. to 200 ft. The Pratt truss is also used as a deck
truss. Most of the Pratt trusses were built in the 1890's to 1920's.

Features

The Pratt truss is a unique truss bearing the name of its originators, Thomas and Caleb Pratt.

The interior diagonal members are all tension members and commonly consist of two or more
eyebars and are fracture critical members. If the diagonal consists of more than two eyebars, it has
internal redundancy. The primary diagonal members all are orientated from top to bottom pointing
towards the truss bottom-center. The single diagonal eyebars orientated from bottom to top towards
the end, are called counters. Counters are tension members which come into play when loads on the
span are in certain positions.

The end diagonals, or end posts are compression members.

The first interior vertical member is commonly called a “hip hanger” as it connects to the hip joint
on the span. This member is a tension member highly prone to fatigue damage and may be an H-
shape which can be pinned or rigidly connected to the truss top connection. The remaining interior
vertical members are generally all box members as they are compression posts, usually pinned at
both top and bottom.

The top chords are compression members and are generally composed of an open bottom box
member with a solid top cover plate and a laced bottom. The bottom chords other than the end two
panels are tension members usually of two or more eyebars. The bottom chord end two panels are
usually a box member and although they are tension members, due to braking and the possibility of
compression in the end two panels, they are shaped to carry compress loads.

The Pratt truss has both top and bottom bracing systems and usually a conventional floor system.
Most Pratt trusses have all truss connections pinned, except that some hanger members may be
rigidly connected to the bottom chord.

Pratt trusses usually have segmental expansion roller bearings.

Pratt Truss Variations


• A camelback type truss is a Pratt truss which is identified by the top chord being in three
chord segments between hip joints. The outer segments slope upward toward mid-span
resulting in greater depth and lower chord stresses in the center of the span.

• A Parker type truss is a Pratt truss of multiple segments of the same section properties
arranged in an arc.

Nomenclature

The standard nomenclature for truss spans, as adopted by AREMA is given in the “Nomenclature”
section of this Handbook. Specific railroads may use varying number systems, such as stringers
numbered from left-to-right or right-to-left, and tower piers from top-to-bottom or bottom-to-top,
and so forth. Many inspectors have slang names for a number of bridge elements which sometimes
makes it difficult in communicating a field problem to the office. Inspectors are encouraged to use
standard nomenclature.

Pratt Truss Deterioration – Scenario

It has been observed that there is a typical scenario for the deterioration of pin-connected Pratt
trusses. At first, there is light wear at the pin joints typified by a rust halo between the pin and
connecting members. As the wear continues, primarily in the vertical compression members at the
pin joints, there is actually a small change in the geometric distance between the top and bottom
chords which results in a change in length for the diagonal eyebars. This change allows the eyebars
to become slack. As the pin hole wear continues to increase in the top chord and in the vertical
members, when a live load crosses the bridge, the floor system deflects excessively under each
wheel load and rebounds between wheels, resulting in flexing of the stringer-floorbeam joints and
causing cracks to form in the connection angels. By the time there is significant wear in the pin
joints (3/8 in. to ½ in. gaps or more), the truss hip plate flexes and retrofit patch plates over the joint
crack shortly after being installed. The sway from upper connection cracks in a fatigue pattern due
to flexing, with the movement coming from the connection being attached to the top chord and
separately to the vertical with movement between the two members.

With extreme wear in the pin holes, the truss becomes quite vulnerable to lateral sway and speed
restrictions become necessary. At the top chord, vertical movement is noted in the chords as the
wheel loads pull downward the verticals. The appearance is like a wave acting at the speed of the
train.

With extensive wear in the pin holes, there is also concern for the pins that will have stress
concentrations at the grooves worn into the pins and the possibility of fracture. Because the Pratt
trusses are generally designed in the E-50 range and comprised primarily of eyebar members, it is
usually not economical to rehabilitate these trusses since it is difficult to add eyebars or otherwise
strengthen the trusses. For those trusses that are not extensively worn, pin plates and other means to
hold the pins in proper location will result in extended span life.

PRATT TRUSS INSPECTION CHECKLIST

Track

Alignment
horizontal on approach
vertical on approach
on structure
surface

Support
behind abutment (hanging ties)
on span

Fastening to deck
movement (sliding)
hook bolts
rail anchors on and off structure

General condition
ties
rails
joints
inner guard rails (if used)
outer guard timbers
spacer straps
spacer blocks

Walkways/Platforms

Walkway surface
holes
tripping hazards
toeboards
fasteners, support members, decking (planks/grating)

Handrail
posts
rails
cables
fastenings

Overall safety items

Truss

Top Chords
worn pin holes in web at verticals
cracked cover plate at hip joint
deteriorated rivet heads on top cover plate
section losses in top cover plate

Bottom Chords
unequally tensioned eyebars
eyebars scored from contracting verticals

End Posts
collision damage from wide loads (compression member)

Interior Diagonals
unequal eyebar tension
worn eyebar heads
accident damage

Counters
loose eyebars
worn eyebar heads
loss of section where striking adjacent eyebars
frozen turnbuckles

Hangers
fatigue crack at top connection
section loss above floorbeam connection
section loss in section at top pin hole
accident damage

Posts
section loss at both ends at pin hole
accident damage

Pins
wear at member contact
grooves at counters and at posts
corrosion under spacer
loose, unchecked thread loose nuts

Top Laterals
section loss adjacent to connection
deteriorated fasteners
loose rods

Bottom Laterals
section loss at end connections, spread back-to-back angles, loss at stringer
connections
connection plates with corrosion losses

Sway Frames/Portal Frames


fatigue cracks in upper connections
damage from high loads

Floor System
see Floor System Section of this chapter for inspection details

Bearings
see Bearing Section of this chapter for inspection details

Emergency

If the inspector discovers a bridge condition that affects the integrity of the bridge under train loads,
contact the railroad dispatcher and responsible authority to stop any trains and arrange for immediate
repairs.

Criteria for the limits of acceptable/tolerable conditions must be provided by the Railroad Owner or
their designated engineer based on the principles of acceptable rating contained in the appropriate
chapter of the AREMA Manual for Railway Engineering.
PRATT TRUSS

Figure 9-155 Pratt Truss – side view Figure 9-156 Pratt Truss – end view

Figure 9-157 Joint U1 – hip joint


Patch plates over crack in bent plat Figure 9-158 Bottom chord – loose eyebars
Joint movement Note bow in bars

Figure 9-159 Pratt Truss – loose diagonal eyebars Figure 9-160 Pratt Truss – loose counter with
section loss
PRATT TRUSS

Figure 9-161 Slippage mark indicates amount to Figure 9-162 Pin and counter wear
wear in joint

Figure 9-163 Lower pin joint Figure 9-164 Interior of upper chord joint
Slippage marks indicate wear in components Note wear at pin and slippage between diagonal
and chord

Figure 9-165 Reinforced stringer – floorbeam


connection due to movement in truss members Figure 9-166 Sway frame connection plate crack
CHAPTER 10
MOVABLE BRIDGES

SECTION 1
INSPECTION

Introduction

Movable bridges are complex structures involving multiple disciplines for their design, maintenance
and inspection. By nature of being movable, these bridges require structural, mechanical and electrical
attention. Often these bridges are located in harsh environments over waterways of either fresh or salt
water, generally they are built close to the water level and therefore are subject to constant moist
conditions, and exposure to marine collisions, all conditions requiring above normal maintenance
needs. Of utmost importance, these bridges must operate reliably.

Movable bridges are constructed over navigable waterways in accordance with U.S. Coast Guard
approved Permits. These Permits specify the required channel width and height that must be maintained
for navigational interests. The requirements for displaying navigational aide lighting at the bridge site
are also given. If the bridge cannot be opened to prescribed clearances, the Coast Guard should be
immediately notified and efforts taken to restore the clearance. Likewise, should the bridge electrical-
mechanical system experience a failure, the Coast Guard needs to be advised of the problem and the
time needed to restore operations.

Early designs of movable bridges were patented by Engineers and thus there were a multitude of types
developed. Many of the more unique types no longer are in service. Today the surviving types of
movable bridges are the swing, bascule and vertical lift. The AREMA Manual for Railway
Engineering, Chapter 15, Part 6 - Movable Bridges provides information concerning specific design
and detail elements of these bridges, both for items common to all types of movable bridges as well as
for items specific for each type bridge.

The format for this Section is divided into general inspection information followed by Operator's
House, Swing Bridge, Bascule Bridge, Vertical Lift Bridge, Electrical Inspection, Mechanical
Inspection, Lubrication, Signal System and Fenders and Dolphins. Within the Section divisions are
included information on track, safety; and, signal and communications.

Inspection

A complete inspection of a movable bridge generally requires a team of inspectors; one for each
discipline - structural, mechanical and electrical. It is important that all inspectors be familiar with the
operations of each type bridge to be inspected and the general functioning of each major component.
Cross-training inspectors are beneficial, particularly for identifying problems that may involve multi-
disciplines.

A movable bridge inspection generally may include all or most of the following elements, depending
on the level of inspection requested:
• System Operational Reliability - all disciplines.
• Structural Inspection - the bridge structural components and supports, special movable bridge
structural elements (for example: bascule segmental girders, counterweights), bridge balance
and seating during operation, the Operator's House, Machinery House, and bridge protective
fenders and dolphins.
• Mechanical Inspection - mechanical components - gearing (open and closed), shafts, couplings,
bearings, trunnions, trunnion bearings, sheaves, pivot bearings, balance wheels, tracks,
counterweight ropes and chains, locks, toggles and wedges, wire rope, brakes, lubrication.
• Electrical Inspection - commercial power, standby power, power transfer switches, motors and
motor control centers, lighting, control system, interlocks, navigation and aerial lights, conduit,
wiring, control desk, panel boards and cabinets, relays, limit switches, indicator lights.
• Signal and Communications Inspection - interlock with electrical system, proper block lights,
radios, telephones, intercom systems.
• Track Inspection – deck conditions, track alignment, miter joints, anchorage against movement
and crowding bridge opening.
• Safety Inspection - operations, lockout during inspections/maintenance, walkways, ladders,
personnel.

One of the most important aspects of the inspection is to discuss with the bridge tender and
maintenance personnel, the operational qualities of the bridge, any changes in the performance of the
bridge, any known problems including those of an intermittent nature and other information. These
personnel have the first-hand knowledge of the bridge. Log books may be kept by the bridge personnel
which may give insight to repetitive problems as well as show the efforts being made in maintaining
the bridge.

The inspection should conduct trial openings to allow the inspector to observe the various functions of
the bridge. Trial openings that have frequent starts and stops lend to better observation of gearing,
bearings and other items. Safety during the operation of the bridge with teams of inspectors must
include procedures for proper activation/de-energizing of the bridge.

Structural Inspection - The structural inspection should follow normal bridge inspection procedures as
described in the Section - Iron and Steel Bridges plus those portions of the chapters on Concrete and
Timber Bridges that apply to movable bridges. In addition, as presented later in this Section there are
specific elements of each type movable bridge that have special structural features. Of primary
importance is the finding of defects, damage, and deterioration to primary elements. Fatigue
considerations and fracture critical members likewise require the same attention as for other bridges.
Certain elements of movable bridges may have complete reversals in stress which should be closely
observed for potential crack development.

Additional structural items include the inspection of control houses, machinery houses, walkways,
ladders, platforms, stairways and elevators. Most movable bridges also have fender and dolphin
systems to protect the site which require close inspections.

Mechanical Inspection - The mechanical inspection is often performed by a mechanical specialist with
years of experience in the proper operations and functioning of mechanical systems and their
components, such as: gearing, shafts, couplings, bearings, brakes, locks, linkages, hydraulic
components, and other component parts of equipment that transmit mechanical power. Observations are
made for alignment, noise, vibration, damage, wear, heat, proper functioning, lubrication and needs for
maintenance or repair.

Most mechanical systems on movable bridges are built for rugged operation and even with some degree
of wear, components generally have long life expectancy. Older mechanical systems often are upgraded
by eliminating long line shafts and installing either hydraulic components to power end items or
electric motors with gear boxes.

Electrical Inspection - The electrical inspection is often performed by an electrical specialist with years
of experience in power and control systems. The inspector should observe the functional operation of
the bridge, look for abnormal performance of the equipment, and identify items that do not conform to
current National Electric Code standards. Items of interest include: incoming commercial power,
standby power sources, power transfer switching, cabling (submarine, conduit and aerial), navigation
lighting, motors, brakes, control panels, and control systems. Newer items requiring inspection are
programmable logic controllers (PLC) computer control systems and the remote control equipment.
Included in inspections should be a review of available spare components at the bridge site. The review
of log book entries indicating faults or other problems should be reviewed with the bridge tender and
others.

Signal System Inspection - The signal system should be inspected to assure that the proper safety
features are in place and working. Of primary interest is the safety link between the electrical and signal
systems, that neither system can be operated without the proper setting of the other system. Other safety
features of primary importance are limit switches assuring that the bridge is in proper alignment and
rails seated.
SECTION 2
OPERATOR'S HOUSE

General

The Control House may be located on the bridge, adjacent to the bridge or at some nearby
designated location.

The Operator's House is the central location for the control of the bridge and related functions.
As such, the Operator needs to have good visibility of the waterway, be able to see the track in
both directions and have easy access to the House.

The inspection of the House includes a visual inspection of all elements of the House - windows,
doors, walls, plumbing, HVAC, lighting and furnishings (desk) for operating the bridge, logging
traffic and centralized location for all means of communications. Since most bridges are operated
on a 24-7 basis with full-time bridge tenders, other furnishings include bathroom facilities and
kitchen facilities (refrigerator, stove, and/or microwave).

In older facilities, a review should be made for operator safety as related to open electrical panels
or moving machinery in near proximity to the operator’s surroundings. Additionally, an alternate
means for escape from the House should be available should there be a house fire. Houses in
some remote locations that are not manned require vandal-proof measures. Bulletproof windows
may be required.

Houses should include a working two-way radio for communications with vessels.

Houses should be equipped with means to clean the exterior side of windows for good visibility
and shades to protect the Operator from sunlight.

Houses should be equipped with alternate signaling equipment, including flags, lanterns, or other
devices.
Operator’s House

Figure 10-1 Operator’s House Figure 10-2 Operator’s House


Metal building on shore Adjacent to bridge

Figure 10-3 Operator’s House Figure 10-4 Operator’s House


1st floor – electrical Excellent visibility
No alternate escape Outside walkway

Figure 10-5 Operator’s House Figure 10-6 Operator’s House


Interior Emergency Replacement
SECTION 3
SWING BRIDGES

General

Swing bridges consist of two-span trusses or continuous girders which rotate horizontally about a
center pivot. Turntables are similar to swing bridges. The swing spans are usually symmetrical
about the pivot but can be unequal in length (bob-tailed) with counterweights provided to
equalize the balance. When in the closed position (ready for rail traffic) all swing span ends are
wedged for support, sufficient to offset any negative reaction (uplift).

The swing bridge offers unlimited vertical clearance but has generally half the horizontal
clearance of a vertical lift bridge of comparable length span. The swing bridge offers the lowest
profile. The swing bridge generally is operated in one direction, or 90 degrees in either direction
or to a skew of the waterway alignment, but some can be positioned end-for-end or for 360
degrees if there are special needs and there are special electrical connections.
Swing bridges are classified by type:

• Center-bearing or pivot support swing.

• Rim-bearing.

The center-bearing type swing spans carry the entire weight of the bridge on a central pivot and
are balanced against tipping by generally 4 to 8 balance wheels. The balance wheels are attached
to the span underside in a circular pattern to ride on a circular track anchored to the pier.
Commonly a circular rack is attached to the track or integral with the track to which drive
pinion(s) engage for movement of the bridge. The center-bearing type span is typically fitted
with six wedge assemblies, two at each end and a pair at the center of the bridge. It is intended
that the end wedges engage the wedge supports sufficient to take the droop out of the span and
raise the ends to a plan elevation. The center wedges stabilize the center and are driven snug so
that they support live loads rather than the pivot bearing but are not intended to lift the swing
span off the center-bearing.

Center-bearing is usually a large casting or housing that holds a multi-stack of bearing elements
called discs or lenses. Traditionally the bearing "sandwich" consists of an upper and a lower steel
disc or lens fixed by pintles to the housing, and with concave inner surfaces to which a central
bronze disc or lens with convex surfaces mate. Oil/grease grooves are cut into the steel lens to
distribute the lubricant uniformly across the entire bearing elements interfaces.

The rim-bearing swing span carries the entire weight of the bridge through a circular girder, to a
ring of tapered rollers (wheels), to a circular track. Their position is held constant by the use of
spider rods connecting the rollers to the center pivot. A center pivot is provided only for keeping
the swing span concentric with the track and circular rack; typically no loads are passed
vertically through the center pivot. The rim-bearing swing span has only four wedge assemblies,
a pair at each end. The bridge live load is transmitted through the end wedges and center ring of
rollers.

Both types of swing bridges have end latch assemblies to assist in proper positioning of the
bridge in the closed position. A positioning variation for swing bridges that rotate in only one
direction is the use of an end bumper for alignment. Swing bridges with center fenders may also
be fitted with either wedge supports or a bumper for the bridge to swing against to hold the open
position.

For both types of swing bridges there is a circular track for the wheels/rollers to move on as the
bridge is operated. A circular rack is attached or often integral with the track for the bridge.
Typically, span mounted machinery drives a pinion(s) that engages the rack to power the bridge
open and closed. Depending on the size of the swing bridge there may be as few as one drive
pinion needed for turning the bridge or as many as four drive pinions. For the multiple drive
pinion arrangement, machinery is provided to assure that the load is shared among the multi-
pinions.

Operational Sequence

To close the bridge from an open position:

1. Sound the horn to close the bridge, assure it is safe to operate the bridge, no marine
vessels approaching, and all maintenance workers clear per railroad safety
procedures.
2. Swing the bridge closed: electrical/mechanical system controls/powers the drive
pinion until the swing span is closed. With the bridge barely moving, the end latch
raises up a ramp and drops into a slot to assist in proper span alignment. With the
bridge in the closed position, the center navigation light facing the marine channel
is showing red to mariners.
3. The drive system is then used to engage the wedges while the miter rails are
lowered into their mating components.
4. The drive system is deactivated.
5. The signal system is activated and if all functions are properly complete, the signal
light will turn to "green".

To open the bridge from a closed position:

1. Reverse process - the navigation lights facing mariners will be green when
complete.

Special Inspection Items

Structural - Alignment is a critical item for the proper operation of a swing bridge and often is
one of the more bothersome items for this type movable bridge. Vertical alignment - the wedges
must be properly engaged or there will be far end lifting when a train first enters the near end.
Alignment can be a problem from a span being heated by the sun on one side, causing warpage
or wear of the bearing surfaces of the wedge assemblies. Other alignment problems are caused
from wear in the pivot or balance wheels which results in the bridge not tracking properly.

Defects can be any of the usual type defects found on fixed bridges such as loss of
section of structural members, fatigue, and wear. Often cracks and loose fasteners are
noted on structural elements in near proximity of the miter rail joints which can emit
excessive vibration and impact from passing rail cars.

Swing bridges feature unique structural systems for transferring loads from the main structural
members (girders or trusses) to the supporting pivot machinery. Typically a pivot girder connects
the girders or trusses to the top of the pivot bearing assembly for center-bearing type swing
bridges. This may be supplemented by a rim girder to support balance wheels. For rim bearing
type swing bridges, a heavy circular rim girder transfers the main structural members’ (girders or
trusses) loads to the roller assemblies. In either case, this supporting structure must be in good
condition, free of major section loss, warpage, or other deterioration which would impact bridge
structural capacity, alignment and operation of the pivot machinery and/or wedges.

Mechanical - Drive system – includes shafts, bearings, couplings, and gear sets. If the bridge
does not track concentrically, the drive pinion will over engage the rack and barely engage the
rack at different points.

Latch assembly - includes a vertical latch bar which upon closing the bridge rides up a
ramp and drops into a slot to position the bridge. Mechanical linkage for lifting the latch
is common with the miter rail lifts and wedge retraction.

Span end machinery - Powered from the longitudinal drive shaft, the span end machinery
usually includes a worm/worm box assembly for powering the primary transversely
orientated drive shaft. Through linkages, a multi-series of parallel shafts supported on
their ends by pillow blocks, operate the latch, rail lifts and wedge assemblies. These
items all receive extreme punishment and wear from vibration in the miter rails with each
passing car.

Center-bearing Pivot - the bearing discs or lenses, housed in the pivot casting are to be in
a bath of lubricant. Generally the pivot contains three discs or lenses - outer steel concave
discs or lenses separated by a convex bronze disc or lens. Without proper lubricant and if
fouled by dirt, the discs or lenses can gaul and even freeze, preventing movement of the
bridge. The bearing elements must be cleaned and relubricated if the pivot should be
underwater in a flood. Pivot bearing wear may be noted at the bearing housing, balance
wheel clearance changes or differences in alignment from driving the wedge assemblies.

Pivot Shaft Bearings – these bearings generally wear very quickly due to the high radial
loads at this location. Reversal of stresses often leads to failure of the bearing cap
fasteners. These bearings additionally are usually in a difficult location for lubricating.

Wheels - Balance and rim wheels must be checked for proper functioning with the
balance wheels only slightly clearing the track while each rim wheel should be in full
contact with the track. Wheel linings wear and may cause center-bearing swing bridges to
dip at the ends upon movement. For rim bearing swing bridges the wheel spider rods hold
the wheels in position and often are found with significant metalwork loss near the center
connection.

Wedge assemblies - include an inclined wedge that should have smooth surfaces, free to
move on lubricated mating surfaces, linkages with pin connections, pivot arms and often
a set of gears for torque reduction. Wedges must be properly engaged to remove the span
end droop and block the center for the passage of live loads. Wedge linkages should be
adjusted so that they are well extended in the driven position to “lock” the wedges in
position.

Electrical – Power, controls and other electrical equipment.

Power - Normal power for movable bridges is from commercial power sources. An
engine-generator set provides an alternate power supply. An automatic transfer switch is
commonly used to switch power sources. Reviewed power system includes the incoming
power, metering, transformers, transfers switches, cabling, termination cabinets, conduits,
lighting and navigational aide displayed lights. Flex cables connecting power on the pivot
pier with the swing span must be checked for wear and any evidence of kinks or pinch
points.

Control System - includes all items for the sequential operation of the bridge. Older
systems may have a drum control for trolley type operation of the bridge while modern
systems use PLC's to computer control all the functions.

Signal & Communications - must function properly for reliable operation of the bridge.
The proper settings of limit switches sensing span closure and rails seated is a must
before the signal system functions can be completed. Radio communications with
dispatcher and marine vessels should be checked.

Track - There are a number of miter rail styles in use including those with a center
section that raises, those with a retracting center hump rail, and those with a center hump
rail that raises. All joints have some degree of difficulty in maintaining surface and
anchorage to the supports.

Common Problems

• span not operating concentric with circular track resulting in drive pinion bearing failure
and poor pinion engagement.

• wedge engagement - sticking, difficult to operate, end lifting.

• span warpage.

• rim wheel frozen, cracked.


• spider rod deterioration at center connection plate.

• component deterioration under miter rails.

• broken circular rack teeth at ends of operation from rough bridge operations.

• lens or discs gummed due to poor lubrication/dirt.

• pivot bearing wear causing the span to sit lower – often reducing the needed clearance
between the balance wheels and track.

• excessive wear of pinion shaft bearings or failure of bearing fastener preload.

• fatigue of bearing supports and machinery support fasteners.

• improper balance wheel clearance allowing swing span ends to strike the rest piers during
operation.

• brake actuators set too rapidly to decelerate high inertial load damaging machinery or
fasteners.
Swing Bridge

Figure 10-7 Swing span – truss Figure 10-8 Swing span – truss
Closed position Operator/Control house with
good vision of channel

Figure 10-9 Swing span – girder Figure 10-10 Swing span – girder
Closed position Open position

Figure 10-12 Swing span – rim support


Figure 10-11 Turntable (Load thru all wheels)
Swing Bridge

Figure 10-13 Rim support Figure 10-14 Rim support


New assembly of wheels Spider rod section loss at center
Spider rods and center plate connection plate

Figure 10-15 Rim support Figure 10-16 Rim support


Arrow – crack in wheel Rack/Drive Pinion good alignment
Note warp in upper track Wheel protruding

Figure 10-18 Journal bearings (top and


Figure 10-17 Poor pinion alignment bottom)
Support for drive pinion assembly
Swing Bridge

Figure 10-19 Pivot support Figure 10-20 Pivot support


Pivot assembly (large bridge) Pivot assembly (small bridge)

Figure 10-21 Pivot support Figure 10-22 Pivot support


Bronze lens with score marks Steel – bronze – steel lenses

Figure 10-23 Pivot support Figure 10-24 Journal bearings (top and
Balance wheel with excessive bushing wear bottom)
Support for drive pinion assembly
Swing Bridge

Figure 10-25 End latch assembly


Centers the span on closure Figure 10-26 End wedge assembly
Arrow – worn wheel

Figure 10-27 End wedge assembly Figure 10-28 End wedge assembly
Converted using electric motor and gear box Converted using hydraulic piston

Figure 10-29 Three –piece miter rail Figure 10-30 Three-piece miter rail
Raised for span movement Seated and locked
SECTION 4
BASCULE BRIDGES

General

The bascule bridge is a movable bridge consisting of either a truss or girder span that is rotated
vertically about a horizontal axis. The span may be rotated on a rolling segment, pivoted and pulled to
a vertical position through linkages or pivoted about a fixed point. The bascule span is
counterbalanced such that the span is approximately balanced in all positions of travel. The bascule
span generally rotates through about 85 degrees. In the lowered or closed position (ready for rail
traffic) the bridge is locked by span locks and for safety, rails are often locked. In the operation of the
bridge, the electrical/mechanical system normally has a drive pinion that engages either a stationary
rack or a stationary pinion that is engaged by a moving rack.

The bascule movable type bridge, particularly the rolling lift type, offers a nearly full open vertical
clearance (85 degrees) and a good length horizontal opening, but typically less than that provided by
a vertical lift bridge. It is primarily used where the channel is not very wide and where vertical height
of an open bridge is a significant concern.

There are three common type bascule bridges:

- rolling lift bascule (Scherzer type)


- multi-trunnion, folding parallelogram (Strauss type)
- simple trunnion (Chicago type)

The rolling lift bascule bridge, developed by Scherzer, rolls back on a circular segment girder atop a
track girder. It is kept in alignment through the engagement of pintles. The advantage of this type
bascule is that as it rolls back, it moves away from exposure to the channel. The Scherzer bascule
bridge has a characteristic vertical counterweight in the air above track clearance that rotates to the
horizontal position just above the track when the bridge is fully opened. The rolling bascule
machinery normally is mounted on the movable span at the center of the rolling radii and the drive
pinion engages a stationary rack mounted on an independent stationary frame. Air buffers are
mounted on the span end to cushion the span as it comes to rest on the bearing pads and span locks
secure the bridge in the down position. The bridge is balanced so that the span is slightly toe heavy
when seated to give the span a positive end reaction and to prevent end flutter or uplift as live load
passes over the span.

The multi-trunnion Strauss bascule consists of a four-sided folding parallelogram bounded by


trunnion joints at its corners. The tower front inclined leg is one side of the parallelogram and is
stationary. The bascule span heel is jointed to the tower base portion of the inclined leg. An overhead
truss which pivots on a trunnion joint atop the tower inclined leg, balances the span to one side and a
counterweight to the other side. An operating strut with an attached rack connects to the hip joint of
the span and pulls to fold the parallelogram to raise the bridge and extends the parallelogram to lower
the bridge. A characteristic feature of the Strauss trunnion bridge is the swinging of the counterweight
from vertical to near horizontal, tucked into the tower. The bridge is balanced so that the toe end is
slightly span heavy to give the span a positive end reaction and to prevent end uplift under the
passage of live loads. The Strauss bascule also has air buffers to cushion seating and span locks.

The simple trunnion bascule commonly called the Chicago type bascule bridge, pivots on a line of
stationary supported bearings. To the rear of the bearing pivot is a counterweight which rotates
downward as the bridge is raised. A rack is attached to the bascule span with stationary machinery
usually mounted on the pier. This type bascule span typically has live load bearings ahead of the
trunnions and rear or tail locks to block live loads to the rear side. Air buffers and span locks are also
used on this type bascule.

There remain several Abt type bascule bridges in service. With the bridge in an open position the
counterweight hangs down from an A-frame tower top bearing. As the bascule span is lowered, the
counterweight rotates and the entire machinery floor rolls up the rear tower legs on a track while
engaging pinions on an adjacent stationary rack.

Although there remain several double-leaf bascule railroad bascule bridges, due to the enormous size
shear locks between leafs and other problems in the transfer of live loads, double-leaf railroad bridges
are no longer in favor. The inspection of double-leaf bascules is the same as single-leaf bascule spans
with special attention to the shear locks or compression/tension connections at the center.

In all bascule bridges, the track structure rotates from horizontal to nearly vertical, thus the need for
very good anchorage to the floor system. Bascule bridges that remain in the open position for long
periods of time may experience gravitation of the track resulting in crowding of the heel rail joints.

Operational Sequence

To close the bridge from an open position:

1. Sound the horn to close the bridge, assure it is safe to operate the bridge; no marine vessels
approaching, all maintenance workers clear, per railroad safety procedures.
2. Lower the bridge: activate the electrical/mechanical system; controls power the drive pinions
which engage the racks. With the bridge nearly closed, the lowering speed is reduced to barely
moving. The air buffers assist in a soft landing while side guides assist in centering the span.
Upon full seating, the span locks are engaged to assure the bridge is seated properly. The
navigation light marking the center of the channel turns from green to red, facing the
mariners. Rail locks are inserted.
3. The drive system is deactivated.
4. The signal system is activated and if all functions are properly complete, the signal light will
turn to "green".

To open the bridge:

1. Reverse process – the navigation lights facing mariners will be green when complete.
Special Inspection Items

Structural

The structural inspection should follow the same format as used for the inspection of any other girder
or truss bridge but adding the inspection of items peculiar to bascule type bridges: segmental girders,
track and treads, pintles, rack frame supports, live load bearings, tail locks, counterweights, bridge
balance and track droop. Bascule bridges by nature of pivoting and tilting allow for the accumulation
of debris and development of corrosion on the floorbeam side facing upward upon bridge opening.
Additionally, if a bascule bridge is normally held in the open position, the weight of the track section
about the weak axis of the floorbeams may show signs of bending.

- Rolling bascule segmental girders are faced with either castings or long curved tread plates.
The casting rarely have problems but the tread plates are known to crack, cold roll the edges
of the girder webs and often have loose fasteners particularly at tread plate joints. Pintle wear
can cause difficulties in proper "tracking" of the bascule as the bridge is lowered.

- The Strauss bascule bridge counterweight truss members adjacent to the counterweight have
a full reversal in loading during each bridge cycle. Fatigue cracks may be found in these
members. These bascule bridges additionally may have a problem with the breakage of
retainer bolts in the trunnion joints and wear between the trunnion sleeve and supporting
gusset plates. Cracks have also been found in the rear hip gusset plates at the interface with
the counterweight concrete. A combined structural-mechanical problem often occurs in this
type bascule bridge in the operating strut guide assembly. The assembly bottom castings have
a tendency to crack adjacent to the drive pinion, wheel bushings wear which affects the mesh
of the pinion with the rack and fasteners are difficult to tighten in the castings/wheel blocks
due to the lack of access.

- Bascule bridges are sensitive to balance. Without proper balance, the motors and machinery
will be overworked and have difficulty in moving the bridge. The correct balanced condition
is for the span to be toe heavy in the lowered position such that the bridge will not try to lift
under live load. Balance can fluctuate during the year due to wet ties, ice or snow load or
extreme dry conditions.

- Bascule bridges with live load bearings commonly have a problem of the bearings being
pounded into the pier concrete and not being effective. The tail locks also are exposed to
heavy impact loads and require regular maintenance.

- The counterweight concrete must be inspected for deterioration as noted in the Concrete
Section. The counterweight pockets for fine adjusting the balance should be closed to weather
as well as to the entrance of birds. Plugged drains within counterweights have been known to
affect balance and in cold regions have caused the cracking of the counterweight concrete
walls.
Mechanical

Most bascule bridges have a similar mechanical system layout for powering the bridge. Commonly
duel motors are located each side and connected to a central gear box or central open gear differential.
From there, line shafts extend outward each side and are terminated with a drive pinion that engages
the rack. Often there is either a secondary gear reduction just prior to the drive pinions or a secondary
gear box. Machinery brakes are usually found on the front drive shafts and to the rear of the motors.
With wear in the mechanical system, it is sometimes found that the two drive pinions do not equally
share the load. This can cause an overloading of gearing and excessive shaft bearing wear. Brakes are
typically poorly covered or not covered resulting in rapid surface corrosion. Left unchecked, surface
corrosion can impair the brakes function.

For older bridges the shafting is supported by journal bearings lined with babbitt or bronze. Older
bridges may have jaw type couplings. Newer mechanical systems use roller bearings and flange, gear
or grid type couplings. Other machinery includes span lock units which are composed of a motor,
gear box and lock bar supported in guides to engage a mating slot. Air buffers may be operated
separately or ganged for common cushioning during bridge seating. Air buffers may experience bent
shafts from the buffer plunger rod extended and as the bridge is lowered the rod drags in an arc on the
pier top until the bridge is seated. Some rods have an end wheel to prevent the rod from being bent.
The rack and pinion behaviors are subject to good bridge alignment.

Electrical – Power, Controls and other electrical equipment

- Normal power usually is from a commercial power source. An engine-generator set provides
an alternate power source. The power system includes incoming power, transformers,
metering, transfer switch, power distribution panels, cabling, conduits, lighting, and
navigational lights.

- Control system includes all items for the sequential operation of the bridge which may have:
resistor banks, contactors, limit switches, cabling, conduits, timers, etc. New systems utilize
PLC's (programmable logic control) for the control of operational functions.

Signal & Communications must function properly for reliable operation of the bridge. Limit switches
for the detection of span lowered, locked and rails seated are generally a minimal check performed.
Radio communications with the dispatcher and marine vessels should be checked for reliability.

Track - The bascule bridge uses a two-piece miter joint located at both ends of the bascule span. The
positioning of the heel miter joint must be such that the joint can rotate without striking the mating
piece as it is raised. The track section on the bascule span must be very well anchored to the
supporting floor system as the track will be rotated to nearly a vertical position. Inadequate rail
anchorage will allow for the rails to slide in the tie plates. On bridges that are held in the raised
position, it is even more important that the track section be adequately secured.
Common Problems

- maintaining bridge balance, holding the toe end down


- span lock engagement tolerance
- rails crowding at the heel of the span
- air buffers, sticking plunger and bent plunger
- the cone gear in the differential gear set is subject to extreme wear and deformation
- journal lining wear
- pinion bearing support fasteners fail
- brake actuators stick
- counterweight deterioration

Scherzer Type
- tracking (worn pintles)
- loose tread plate fasteners
- cracked tread plate connection angles/worn web plates
- corrosion of embedded steel grillages/track/rack frame

Strauss Type
- broken trunnion retainer bolts
- fatigue cracks in counterweight truss
- operating strut guide assembly loose fasteners, poor gear mesh, casting cracks

Chicago Type
- poor live load bearing pads
- poor tail lock conditions
- worn trunnion bearing bushings
Bascule Bridge

Figure 10-31 Scherzer type bascule Figure 10-32 Scherzer type bascule
Rolling bascule span Rolling bascule span
Truss Girder

Figure 10-33 Strauss trunnion bascule Figure 10-34 Strauss trunnion bascule
Lowered position Partially raised
Note member parallelogram

Figure 10-35 ABT type bascule Figure 10-36 ABT type bascule
Lowered position Raised position
Bascule Bridge

Figure 10-37 Rolling bascule Figure 10-38 Rolling bascule


Segment girder Segment girder
Track Tread plate, worn pintles

Figure 10-39 Rolling bascule Figure 10-40 Rolling bascule


Segment girder Floorbeam
Tread plate deflection Debris collects and section loss

Figure 10-42 Rolling bascule


Figure 10-41 Rolling bascule Cracked rack section
Pinion and rack Thru bolt holes
Bascule Bridge

Figure 10-43 Strauss type bascule Figure 10-44 Strauss type bascule
Counterweight trunnion joint Broken retainer bolt
Worn pin-gusset hole Common with joint wear

Figure 10-45 Strauss type bascule Figure 10-46 Strauss type bascule
Counterweight hip connection Operating strut with rack
Crack due to interface corrosion Operating strut guide

Figure 10-47 Bascule span – raised position Figure 10-48 Bascule span
Floorbeam – bowed from track load Toe end miter rail joint
SECTION 5
VERTICAL LIFT BRIDGES

General

Vertical lift bridges consist of a movable truss or girder span between fixed lifting towers. The lift
span is counterbalanced by counterweights located in the towers with wire ropes running over tower
top sheave assemblies connected to the span and counterweight. The manner in lifting the span
determines the type vertical lift bridge. Vertical lift bridges provide a wide horizontal channel
opening but limit vertical clearance. The requirements for the opening are determined by the Coast
Guard.

Vertical Lift Bridge Types

- tower drive
- span drive

The tower drive type vertical lift bridge operates by mechanically turning the tower top sheaves
which move the counterweight ropes attached to the span and counterweight. A complete set of
mechanical drive equipment is located atop each tower. Drive pinions engage circular rack sections
attached to each sheave. Electronically both towers must be synchronized to keep the span level as
the span is raised and lowered. The lack of keeping the span level is called "skew" and can be a
significant problem with this type bridge.

The span drive bridge is operated from the lift span through a central mechanical system that powers
an independent system of wire ropes called "operating ropes or haul cables." There are two types of
span drive bridges; those totally operated by cables and those operated by drive shafts to span end,
operating rope drums. The operating rope ends are attached to the top and bottom of the tower with
the ropes spooled on the drums in several turns that cross the drum in spiral grooves. There are two
advantages to the span drive type vertical lift bridge: one set of machinery and the span is less likely
to skew.

The vertical lift bridge is set to be balanced or in equilibrium when the bridge is in the raised position
and to be span-heavy in the lowered position. This is accomplished by the passage of the
counterweight ropes from one side of the tower sheaves to the other side. When the lift span is down
the counterweight ropes are long on the span side and short on the counterweight side. When the
span is in the up position, the opposite is the case. For bridges with lifts over 40 feet, there usually is
an auxiliary counterweight system or balance chains connected between the bottom of the
counterweight and tower to make further weight adjustments as the span travels, without causing
excessive changes in power requirements.
Operational Sequence

To close the bridge from an open position:

1. Sound the horn to close the bridge, assure it is safe to operate the bridge, no marine vessels
approaching; all maintenance workers are clear, per railroad safety procedures.
2. Vertical lift span lowers while the counterweight rises in tower guides.
3. The channel center navigation light turns from green to red indicating closure.
4. The span slows in speed as it approaches the pier top and is cushioned by air buffers located
at each span corner.
5. The span guides and a span centering pin position the span into correct position so that the
rails properly engage a mating section mounted on the span end.
6. Upon seating, span locks are engaged and the span operating equipment is deactivated.
7. The signal system is activated and if all functions are properly complete, the signal light will
turn from red to green.

To open the bridge from a closed position:

1. Reverse process - the navigation light facing mariners will be green when complete.

Special Inspection Items

Structural

The structural inspection should follow the same format used for the inspection of any other girder or
truss bridge and viaduct towers but adding the inspection of items peculiar to vertical lift bridges:
counterweights, bridge balance, guides (counterweight and span), machinery house or tower
enclosures, counterweight and operating ropes, sheaves, and centering devices.

- The counterweight and operating ropes should be carefully examined for broken wires,
excessive surface wear, tightness and sharing of load, end attachments, and indications of
core collapse. As the ropes age, they elongate and/or stretch. For operating ropes it becomes
necessary to adjust the ends of the ropes which have devices for making these adjustments.
The only adjustments made on counterweight ropes are to adjust for individual loose ropes
which are uncommon after initially being adjusted.

- Older sheaves are castings and while most newer sheaves are made from weldments. A
careful inspection is needed to look for possible crack development in the sheave webs. The
sheave shafts commonly have a reduction in diameter at the interface with the sheave hub.
An in-depth inspection is needed at a determined interval based on age, details and bridge
cycles to ultrasonically look for cracks in the shaft. The shaft should be found tight within the
hub of the sheave. No fretting corrosion should be evident. The sheave grooves should be
checked that there are not conditions present that would nick or score the wire ropes. The
grooves should cradle the ropes.
Mechanical

Commonly the bridge mechanical layout consists of dual motors connected to a central gear box to
which drive shafts are connected either to end pinions that engage the circular racks on the sheaves
(tower drive) or engage wire rope hoist drums. As bridges age there should be a check that there is
equal torque being transmitted to each end item. The mechanical system includes machinery brakes
on the drive shaft and motor brakes. Older bridges have line shaft journals of babbitt or bronze and
newer bridges have roller bearings for supports.

- For vertical lift bridges, additional items for mechanical inspection are guide rollers for the
lift span and the counterweight, sheave journals/roller bearings, drums, idler and deflector
sheaves, air buffers and span locks.

- Lubrication is needed throughout the vertical lift bridge on moving mechanical components,
guide wheels, brakes, air buffers, gear boxes, wire ropes and other items. Each vertical lift
bridge should have a lubrication chart for that bridge and means to access all points of
lubrication with the proper application tools.

Electrical - Power and Controls

- Normal power usually is from a commercial power source. An engine-generator set


provides an alternate power source. The power system includes incoming power,
transformers, metering, power transfer switch, power distribution panels, cabling, conduits,
lighting, navigation lighting and aerial or loop cables between the towers and lift span.

- The control system includes all items for the sequential operation of the bridge which may
have: resistor banks, contactors, limit switches, timers, selsyn transmitters, radio data links,
etc. Newer bridges often use Programmable Logic Control (PLC) systems for the control of
operational functions. These systems can also diagnose system problem areas. For tower
drive bridges, the timing between towers is critical for skew control. This is often a source of
problems.

Signal & Communications must function properly for reliable operation of the bridge. Limit switches
generally are located to detect that the lift span is seated, rails are engaged and rails are locked.
Communications between the control house and the tower top machinery rooms is a safety
consideration.

Track - The vertical lift bridge normally has a two piece miter joint at each end of the bridge. Some
railroads prefer a three piece set without raising the center section. The miter rails and their supports
receive severe punishment from passing wheel loads and require close inspection for cracks, wear
and loose hardware.
Common Problems

- span balance and bounce upon seating


- span skew for tower drive bridges
- loose operating ropes for span drive bridges
- span warpage and uneven support at the span bearings
- air buffer plunger rods stick
- lack of lubrication on counterweight ropes
- counterweight sheave trunnion cracking – particularly at the fillet
- counterweight sheave bearing and drum pinion bearing accelerated wear
- counterweight deterioration
Vertical Lift Bridge

Figure 10-49 Vertical lift – lowered position Figure 10-50 Vertical lift – raised position
Tower drive Tower drive

Figure 10-51 Vertical lift – lowered position Figure 10-52 Vertical lift – raised position
Span drive Span drive

Figure 10-53 Vertical lift – hydraulic lift Figure 10-54 Vertical lift – screw jack lift
Infrequent openings
Vertical Lift Bridge

Figure 10-55 Sheave-bearing assembly Figure 10-56 Sheave-bearing assembly


Tower drive Tower drive
Roller bearings Trunnion bearings

Figure 10-57 Sheave trunnion shaft Figure 10-58 Sheave trunnion


Surface condition inspection Bearing assembly

Figure 10-59 Lift span centering device Figure 10-60 Lower air buffers
Vertical Lift Bridge

Figure 10-61 Counterweight ropes Figure 10-62 Counterweight ropes


Cable splay at counterweight Fiber core being expelled

Figure 10-63 Counterweight ropes Figure 10-64 Counterweight ropes


Broken strand Severe crown wear

Figure 10-65 Counterweight ropes Figure 10-66 Operating ropes-idler wheels


Raised spelter at socket Excessive rope looseness
Normal conditions
Vertical Lift Bridge

Figure 10-67 Span drive – bridge raised Figure 10-68 Counterweight


Hood over sheaves (background) Balance pockets – weights
Machine House and drive shaft (foreground) not uniformly distributed

Figure 10-69 Span lock Figure 10-70 Tower access


Caged ladder and elevator

Figure 10-71 Two-piece miter rail Figure 10-72 Two-piece miter rail and signal
Worn points equipment
SECTION 6
MOVABLE BRIDGE-ELECTRICAL INSPECTION

General

An electrical inspection of the power and control systems of a movable bridge is as important as
any other portion of a bridge inspection. The reliability of the bridge to function upon demand is
imperative. An electrical inspection may be a simple visual inspection of the various components
observing their functioning during bridge operations combined with meter readings. Or an
inspection may be in-depth in which wire insulation is megger tested, motors are opened and
bridge electrical functions are monitored and recorded on strip logs as the bridge is operated
through multiple cycles.

In any electrical inspection, safety measures need to be established before the start to assure that
the bridge power is not accidentally turned on or that the bridge is not moved. Established lock-
out procedures must be followed. Where not established, the inspectors should establish an on-
site procedure or measure as simple as a tag placed over the controls stating inspection-in-
progress may suffice to remind the bridge operator that inspectors are at work.

The inspection should also include a discussion with the bridge operator to learn what he has
observed concerning the electrical system while operating the bridge. A log book of all electrical
maintenance work and any "call-outs" to the bridge for emergency work should be available at
the bridge and reviewed.

Power Supply

The normal power supply, standby power supply, and standby generated power should all be
examined. A power failure of the normal power supply should be simulated by turning off that
power to determine if the standby power will be activated and come on-line.

Items to Inspect -

- power pole, weather head, and transformers


- power transfer switch
- engine-generator set, fuel supply means
- batteries and battery charger
- conduit runs, condition, connections
- termination of wires in cabinets
- wire identifications
- noise, smells and heat build-up
- panel boards, circuit breakers
- abandoned wires
- lights
- HVAC
- lightning protection
- disconnect switches
- submarine cable & terminations
- loop or festoon cables
- flex cables

Control Systems

For all the movable bridges with electrical control systems there are basically three type of
systems: Manual, typically with a drum controller and associated with resistor banks, semi-
automatic, typically with a relay control system, and programmable logic control (PLC) system.
Movable bridges prior to the mid-1960's were all drum control or trolley car type controls where
the operator manually turned a lever to engage various levels of resistance which controlled
speed. The semi-automatic system allowed for push-button automatic sequencing and allowed
for the use of either drum controllers or solid state drives (e.g. Silicon Controlled Rectifier
[SCR]) with sequence and speed controlled by relay logic. By the 1990's PLC systems were the
state of the art in bridge controls. The computer systems presently being used are very reliable
and have diagnostic capabilities which greatly reduce bridge "down-time." PLC systems are
typically used in association with solid state drives such as SCR or AC vector drives.

Each type control system allows manual operation that will enable the bridge operator or the
inspector to operate the bridge item-by-item, i.e. by push-button retract span locks, release
brakes, energize motors, etc. An electrical inspection may simply observe the individual
functioning of items powered by the control system or it may entail a complete logging of the
electrical characteristics of an item during the complete cycling of the bridge.

Items to Inspect -

- motors - amperage draw, vibration, brushes, bearings, lube, heat build-up


- other motors - brakes, lock bars - same inspection items
- limit switches - function, water tight, damage, good mountings
- contactor, relays and switches contacts - loose, vibration or chatter, pitted, stuck
- by-pass switches - usage
- system operates per design logic and all safety interlocks are in place
- signal system interfacing performs as required
- maintenance log book reported problems
- bridge operator's observations
- indicator lights and span position indications
- lightning protection

Remote Control

Remote control is when a person operates the bridge from a remote location, usually from a
location from which the waterway cannot be viewed directly. Such operations have typically
been from dispatcher offices where CCTV monitors display site conditions. The remote
operation is through either microwave tower or land-line linkages. For a bridge to be remote
controlled a Coast Guard permit is required. Not all sites are suitable for remote operations,
particularly those where there are frequent marine collisions, recreational activities around the
bridge and a high volume of marine and rail traffic. Special site considerations are necessary to
assure safe operation of the bridge. Typically such operations require: an electrical system
compatible for such operations, a reliable mechanical system, a compatible signal system, CCTV
system to observe the pier tops and waterway to assure a safe closure, microwave boat detection
for poor visibility conditions, marine radio broadcasting usually at 1 minute intervals for 10
minutes identifying the bridge and warning of the lowering, and indications that the bridge is
lowered and locked in position.
Electrical

Figure 10-73 Typical incoming power line Figure 10-74 Power transfer switch
Automatic function

Figure 10-75 Engine-generator set Figure 10-76 Special electrical cable loop
Backup power supply Vertical lift bridge

Figure 10-77 Control deck Figure 10-78 Control desk


Drum/trolley control Push button control
Electrical

Figure 10-79 Electrical panel board Figure 10-80 Control cabinets


Open wiring Fully enclosed

Figure 10-81 Resistor bank Figure 10-82 Drum control


Located in open area Note series of contacts

Figure 10-83 Limit switch with vertical plunger Figure 10-84 Navigation light
rod
SECTION 7
SIGNAL SYSTEM

General

The signal system is a key component in the operation of a movable bridge. Without the
proper interaction between the signal system and the bridge electrical system, the bridge
will not operate safely. The signal system on a movable bridge, besides providing
continuity within a given block, also detects that the bridge is in proper position for the
passage of trains and that the rails are in proper alignment tolerance.

The interaction between the signal system and bridge electrical and structural systems at
times becomes a problem in achieving the required tolerances as established by FRA for
bridge seating and may require a joint inspection by signal and B&B personnel. Besides
worn and misadjusted signal linkages, ties under miter joints, bridge balance, worn span
locks, and other items may contribute in not achieving proper signal tolerances.

The inspection of the signal system is normally performed by the Railroad’s Signal
Department and generally performed independent of a normal bridge inspection.
Signal

Figure 10-85 Signal control console Figure 10-86 Signal control display panel
over control console

Figure 10-87 Signal contract points at miter Figure 10-88 Signal control boxes at
rail joint miter rail joint

Figure 10-90 Signal control boxes for


Figure 10-89 Signal control display panel miter rail seated confirmation
SECTION 8
MECHANICAL INSPECTION

General

A mechanical inspection of a movable bridge is as important as any other portion of the bridge
inspection. The reliability of the bridge to function as designed is imperative. A mechanical
inspection may be as simple as making visual observations of the system both in a static
condition and while being operated or as complex as opening journal bearings and gear boxes
and strain gauging the bridge for proper balance through recording shaft torque.

A mechanical inspection must start with a clear understanding between the bridge operator and
the inspectors on safety policies that will be in effect during the inspection period. It is prudent to
open the main circuit breaker if possible to prevent accidental operation of the bridge and have a
sign or tag placed over the controls reminding the operator of the inspection in progress.

The inspection should include a review of maintenance log books and a discussion with the
bridge operator concerning problem areas and deficiencies noted. Typical signs of problems
include vibration or undue heat generation.

There are a lot of similarities between the mechanical systems for various type movable bridges.
All of the bridges have open gearing and/or closed gear boxes; journal bearings and/or roller
bearings; couplings; shafts with keyways, keys, and collars; brakes and lubrication. Specialty
items include air buffers, span locks, balance and rim wheels, tracks, racks, latches, guides,
wedges, sheaves, trunnions, machinery supports, and pivots.

Items to Inspect

- open gears - alignment, mesh, wear, tooth breakage, lubrication


- closed gear box - contaminated lubricant, gear wear and pitting, metal particles,
bearings
- journal bearings - lining wear, galling, worn or packed grease grooves, shaft
smoothness, housing, supporting group, and housing anchorage, lubrication, general
alignment
- roller bearings - roller and raceway surfaces, contamination, lubrication, seals, housings
- couplings - alignment of mating parts, tight, proper support, seals, lubrication
- shafts - operate without warpage, distortion or cracks in keyways, losses from corrosion
- collars - properly positioned and tight on shafts to hold shafts in position
- brakes - drum surface true and not glazed, brake pad thickness, pads fit drum, release
and set correctly, proper torque settings, fluid levels, alignment of pads to drums (discs)
- air buffers - clean and smooth cylinder and plunger rod, functions (pressure), free drop,
tolerances
- span locks - bar engages slots, gear box conditions, controlled by limit switches
- balance, rim, guide wheels - rolling surfaces, axle linings, lubrication, alignment - track
- proper alignment and anchorage, cracks, warpage
- racks - alignment, mesh, wear, tooth breakage, lubrication
- latches - linkage, roller wheel, alignment
- wedges - smooth surfaces, pins, linkage, settings, alignment, lubrication
- pivot - lens or disc wear, surface smoothness, housing interior clearances, lubrication

Hydraulic Operations

It has been found in some situations that in rehabilitation of movable bridges, hydraulic
systems can been cost effective for total mechanical system replacement. Such a system
usually includes: a “shelf packaged” hydraulic power plant that is electrically operated
with possible small diesel generator back-up, piping to hydraulic units to control end
items, hosed and cylinders. Such installations may directly power drive pinions, actuate
brakes, and cylinders control wedges and rail lifts.

Maintenance includes checking and changing hydraulic filters, observing connections for
leaks, hose deterioration, and changing fluid at programmed times.
Mechanical

Figure 10-91 Bascule bridge Figure 10-92 Closed gear box


Open gearing – differential Inspection port on top
Poor cone gear circled

Figure 10-93 Gear-pinion set Figure 10-94 Gear-pinion set


Poor alignment and worn teeth/plastic flow Poor lateral alignment

Figure 10-95 Bearing with Babbitt lining Figure 10-96 Bearing base with deteriorated
Lining worn and shaft with movement/thrust Babbitt lining
Mechanical

Figure 10-97 Split journal bearing Figure 10-98 Journal top lining
Shaft highly polished Good condition, needs cleaning;
grease groove

Figure 10-99 Jaw coupling Figure 10-100 Grid coupling


Poorly aligned Good condition

Figure 10-101 Flanged coupling Figure 10-102 Shaft keyway


Distorted and cracked
Mechanical

Figure 10-103 Open gear box


Worm gearing Figure 10-104 Beveled gear set
Collars #1 and #2 prevent thrust

Figure 10-105 Motor brake Figure 10-106 Machinery brake

Figure 10-107 Disk Brake Figure 10-108 Interior of closed gear box
SECTION 9
LUBRICATION

General

All movable bridges require lubrication of the right type lubrication and applied at the
correct frequency. A number of newer bridges have lubrication charts located in the
machinery room or control house where the bridge maintainer can review the point of
lubrication and the type and frequency of applying. There is no universal lubricant for all
components of a movable bridge.

The determination of proper lubricants should be performed by a lubricant specialist who


will review the particular application for pressures, speed and temperatures. The
specialist can also advise on the amount and frequency of lubricant application.

A relationship has generally been found between items properly lubricated and access to
that location. If access is not readily available, more than likely the item is not well
lubricated. For some bridges with poor access, extensions from grease ports to the nearest
access point have been established or in other cases, automatic lubrication systems or
pressure grease cups have been installed.

Items to Inspect

An inspection of a movable bridge should include the inspection of all lubricated and
hydraulic items. The inspection should note whether the lubricant is effective, amount is
adequate and if the lubricant is contaminated.

Common items requiring lubricants –

Swing Bridges - pivot bearing, line shaft bearings, differential, clutch, worm
gear box, linkages, wedges, circular rack and track, balance and rim-bearing wheels, open
gearing/enclosed gear boxes, hydraulic brakes, motor bearings, latch assembly parts, and
other items as they may apply.

Bascule Bridges – line shaft bearings, differential, racks and pinions, span locks,
centering devices, hydraulic brakes, motor bearings, open gearing/enclosed gear boxes,
track and tread plates, pintles, air buffers, and other items as they may apply.

Vertical Lift Bridges – line shaft bearings, differential, span locks, centering
devices, hydraulic brakes, motor bearings, open gearing/enclosed gear boxes, guide
wheels and guide contact surfaces, operating and counterweight ropes, sheave and idler
wheel bearings, air buffers, and other items as they may apply.

Standby engine-generators require an inspection of lubricants and fluids. If compressed


air is used on a bridge, the air compressor also should be inspected.
During an inspection, should there be any concern for the quality of the lubricant,
particularly in enclosed gear boxes, a sample of the lubricant should be taken and
analyzed. Such testing can be performed by the same lab that analyzes locomotive engine
oil. Oils should be replaced on set intervals even if testing does not reveal deficiencies.

Fluids and filters for hydraulic drives on movable bridges likewise should be routinely
inspected.
Lubrication

Figure 10-109 Automatic lube system Figure 10-110 Automatic lube system
Journal bearing Journal bearing

Figure 10-111 Journal bearing Figure 10-112 Journal bearing


Lubrication box Excessive lube

Figure 10-113 Roller bearing Figure 10-114 Lubrication chart


Packed with lube
SECTION 10
FENDERS AND DOLPHINS

General

The inspection of fenders and dolphins protecting a movable bridge should be a part of
the bridge inspection. The inspector should view the protection system from two
standpoints – do the fendering and dolphins protect the bridge adequately and, do any
components of the fenders and dolphins pose a threat to the safety of passing marine
vessels. The answer to adequacy of protection may require an analysis based on the type
and size of vessels using the waterway. Such methods of analyses are available in the
AASHTO Manual published by FHWA.

There are many types of protection systems and some are featured in the AREMA
Manual for Railway Engineering, Chapter 8, Part 23 – Pier Protection Systems at Spans
over Navigable Streams.

The field inspection of the protection system should look for loose and/or missing
components such as timber rubbing strips or wales, exposed bare steel supports which
could damage a vessel, and deteriorated components including fasteners and wraps that
could not withstand a vessel impact. Any infringement of the protection system into the
authorized waterway channel should be promptly corrected.

The inspection should include a review of adequate access and functioning of


navigational aids such as the required channel light markers and air clearance gages
where required by the Coast Guard.
Fender and Dolphin

Figure 10-115 Fender system Figure 10-116 Fender system


Lining the channel Lining the channel

Figure 10-117 Timber fendering/dolphin Figure 10-118 Fender/dolphin system

Figure 10-119 Dolphins without fender Figure 10-120 Typical dolphin


CHAPTER 11
TUNNEL INSPECTION

General

Tunnels present unique inspection challenges. Access for inspection is very track time sensitive.
A good lighting source is required to properly evaluate tunnel condition. Tunnels are
constructed with many techniques and materials, and it is not uncommon to find more than one
of these construction types within one tunnel. Depending on the type and condition of the
tunnel, considerable expertise may be required to properly evaluate the structure.

Safety

Inspections conducted through tunnels require extra safety precautions. For confined space
issues, refer to Chapter 2, Confined Spaces of this Handbook. The inspector must comply with
all appropriate Federal, State, Provincial, and local laws as well as Railroad Operating Rules.

The lack of light throughout a tunnel shaft makes it a mandatory requirement for the inspector to
be prepared by having access to a good bright light to provide adequate illumination for
inspection purposes. Much of the equipment required for bridge inspection is also appropriate
and necessary for tunnel inspection. Lists of that equipment are given in Chapter 1, Bridge
Inspection Practices of this Handbook.

It is not uncommon for local wildlife to be discovered in tunnels, either living within the tunnels,
or using the tunnels as transportation corridors. That wildlife includes snakes and a variety of
mammals, insects, and bats. The inspectors need to be aware of these inhabitants while entering
and inspecting the tunnel.

Emergency inspections due to fires, floods, earthquakes, and derailments present additional
safety hazards and special safety precautions may be needed.

Tunnel Inspections

External Environment

The inspector needs to be aware of the outside environment over and around the portal areas, and
tunnel shaft. The inspector needs to look and make note of:

• Vegetation: Extensive vegetation and/or tree growth near the tunnel portals that may
impact the stability of the tunnel portal areas.

• Ground Water: Note any seeping ground water or any indications of ground water. If
ground water is apparent, note how and where the water is draining.
• Adjacent Slopes: Slopes adjacent to, above, and near the tunnel portal for signs of slope
instability and any impending slides. See Figure 11-1.

• Rock Slopes: In a rocky environment, look for new fault lines, cracks or movement of
rocks, especially directly above or adjacent to the tunnel portals. Note any recent rock
falls near the tunnel portals. See Figure 11-2.

Internal Tunnel Safety

Older, longer tunnels were usually constructed with motor car setouts or man-size cutouts along
the side of the tunnel. The inspector needs to be aware of these areas and inspect them. Some
tunnels may have safety items and utilities such as fire alarms, call boxes, lights, and fiber optic
conduits. If the tunnel has these types of miscellaneous additions, the inspector needs to note
any obvious damage or deficiencies.

Drainage

Inadequate or non-functioning drainage is one of the biggest causes of problems for all
components of a tunnel. Most tunnel shafts were constructed deep enough in the natural ground
to experience ground water problems. The inspector needs to observe and note all free water
conditions in and around the tunnel, including:

• Inspect for free flow access of storm or ground water from one end of the tunnel to the
other to ensure water is not trapped within the tunnel shaft. See Figure 11-3.

• The tunnel shaft lining will either need to completely restrain all the ground water, or
there needs to be a free flow of the ground water into a controlled channel for discharge
out of one or both ends of the tunnel. See Figure 11-4.

• Record the locations of any ice accumulations.

There are three main components of a tunnel inspection and each of these components, although
unique to a specific tunnel, exists at every tunnel location. These components include the
Portals; Shaft; and the Environment. A complete tunnel inspection should include a review of
each of these components.

Environment (Gas)

Just as most tunnel shafts were constructed deep enough in the natural ground to experience
ground water problems, it is also common for the tunnel shaft to have opened a natural gas
pocket. The owner is often aware of these conditions with this information passed down
throughout the years as well as noted in previous inspection records. For this reason, the
inspector needs to review previous inspection records before starting a new inspection.
Portals

Tunnel portals are the openings at each end of the tunnel. Portals usually include wingwalls and
a headwall, which are typically constructed of concrete, timber, or stone masonry. There are a
few tunnels that have been constructed in very stable geological strata and may not require
wingwalls or a headwall to support the surrounding ground. See Figure 11-5.

If a tunnel does not have any wingwalls or headwall, the inspector should review ground
conditions around the portals to verify that stability is maintained. The review should also verify
that there are no signs of current or potential ground slides that will cause debris to fall onto the
track or in front of the tunnel causing a blockage of the opening and foul the track. These
observations should include a review of general surface drainage patterns to verify that
concentrated flows are diverted around the portals, to the extent possible. See Figures 11-6 & 11-
7.

Portal Headwalls are located over the top of the tunnel opening. Because there was usually a soil
cover that the tunnel would need to penetrate to get to stable ground or rock, and to shorten the
tunnel, a headwall needed to be constructed, much like a retaining wall, to support the
overburden above the tunnel shaft. Most tunnel headwalls have the date the tunnel was
constructed stamped/cast into the concrete or stone. If through the years, the tunnel was
extended, the original date of construction along with the modification is often shown in the
headwall. Inspection of the headwall would include the following items: See Figure 11-8.

• Verify the headwall is sufficient to retain the ground without any debris, such as
boulders, stones, trees, etc., falling over the top of the wall onto a train or fouling the
track. Similarly, verify that the headwall height is adequate to divert drainage around the
portal, and that the drainage channels are not blocked.

• The stability of the headwall needs to be inspected to verify there are no lateral
movements at either the top or bottom of the wall. Excessive seepage through the wall
could be indicative of high ground water pressure and could lead to accelerated
deterioration of the headwall materials. See Figure 11-9.

• Verify the structural integrity of the headwall with consideration for the type of material
used for construction.

• For a timber headwall, check for decay as well as obvious failures such as cracks or gaps
between members. See Figure 11-11.

• For concrete headwalls check for any cracks, deterioration, or movement at joints.

• For stone masonry, check for slippage between joints, deteriorated grout, or cracked and
loose stones.

Portal Wingwalls are constructed on one, or both, sides of the portal entrance, extending from the
headwall to the footing below ground. However, it is not unusual for a tunnel portal to have a
headwall and not have wingwalls if the adjacent ground is stable along each side of the track as it
exits the tunnel. The function of the wingwalls is the same as a retaining wall. Inspection of the
wingwalls would include the same items as for portal headwalls listed above. Because wingwalls
are usually supporting the side of a mountain or large hill, it is likely that there will be water in the
ground behind the structure. Check the weep holes in the wingwalls, if present, to make sure they
are working properly. If there are not weep holes, check for signs of excessive water pressure or
seepage to determine if weep holes need to be installed. See Figure 11-10.

Tunnel Shaft

The tunnel shaft is the fundamental tunnel structure supporting the opening from one end of the
tunnel to the other. There are tunnels that have been constructed through very stable rock material
that have never required support. Other than the exposed natural rock, there would be no other
structural support through the tunnel with the possible exception of rock bolts or anchors in the
roof or walls to stabilize local rock slabs or loose wedges. See Figures 11-12 & 11-13.

Many tunnel shafts are constructed through material that will require support to keep the shaft
stable. The extent of the supports is proportional to the stability of the ground. The following
discussion describes typical types of liners that can be found in railroad tunnels:

• Timber Sets: Timber sets are fabricated with large posts and consist of a column on each
side of the tunnel and a top arch connecting the two sides. The arch is fabricated from
pieces of material of the same size, and cut at the ends at angles so that, when put
together, the pieces will form an arch. The base of each of the vertical columns may be
supported on a mud block, a slab of concrete, or a continuous bottom sill that runs along
each side of the tunnel. The timber sets are placed at spacings determined by the ground
stability.

• Steel Sets: Steel sets are fabricated and installed in the same manner as the timber sets.
Steel sets are found in the newer tunnels, or have replaced timber sets because of tunnel
modifications. Spacing of the steel sets will vary for the same reason as the timber sets.
See Figure 11-20.

• Lagging: Lagging material, such as timber or steel sections, is placed horizontally and
behind or between timber or steel sets to support unstable ground. The lagging can be
installed the full height on both sides of the tunnel, as well as over the top of the arches,
making a solid structural tunnel shaft lining. Lagging is installed where ground or rock
stability problems within the tunnel are located. See Figure 11-21.

• Cribbing or packing: Cribbing or packing can be found behind sections of lagging and
will consist of anything, from various sizes and lengths of timber to pieces of stone or
concrete. At sometime in the tunnel’s history, a section of the tunnel shaft may have
failed. The entire void was filled with cribbing or packing between the failed ground
and lagging. See Figure 11-14.
• Concrete or masonry lining: When the stability of the tunnel shaft becomes a serious
issue, a solid plain or reinforced concrete or masonry wall and top arch will be
constructed to support the opening. It is not uncommon for the tunnel shaft to be cut
through a geological fault plane that will require the strength of a reinforced concrete
wall to maintain tunnel stability. In older tunnels these conditions were resolved using
masonry lining such as stone or brick. See Figures 11-15 & 11-16.

• Steel lining: Corrugated steel plates can be used as tunnel liners. Consideration for
corrosion must be considered due to ground water conditions.

• Tunnel Floor: Many tunnel floors remain as the natural ground that existed when the
tunnel shaft was cut. The track structure is placed on the natural ground surface. For
those tunnel shafts with a more serious stability problem, struts may be found placed
below the ballast section from one side of the tunnel to the other. For those tunnels with
serious stability issues, the tunnel floor may be constructed as a solid reinforced concrete
section.

For each type of tunnel shaft construction the inspector needs to have an appreciation of why each
section needed to be installed. With this information, the inspector will have a better
understanding of what types of problems to watch for and why they occur. Good inspection notes
must be maintained and be available for future inspections. It is always a good idea to refer to any
previous notes so that any changes in conditions can be identified. Inspection of the tunnel shaft
should include the following items:

• Review the general condition of any natural or exposed material through the length of the
tunnel shaft. Look for signs of cracks, openings, or movement. For unlined (exposed
bedrock) tunnels look at the floor for new pieces of rock that may have fallen and review
the roof and wall areas for clues of fresh material indicating that a section has broken
away. Where rock bolts or anchors have been installed in the arches and crown areas, look
for loose bearing plates, excessive corrosion of steel components, and fresh rock falls or
slabbing around bolt anchor locations. See Figures 11-18 & 11-19.

• Timber Sets: Look for signs of visual decay as well as “sounding” and drilling suspect
areas of the timber to test for voids. Locations for decay will be along the bottom sill, the
bearing area at the bottom of the columns, or at any bolt or drift pin locations. An
important clue when inspecting timber sets will be looking at the joints of the top arch for
any openings. Watch for any other signs of movement or stability issues.

• Steel Sets: Inspect for corrosion, section loss, and structural problems such as movement
along the base of the columns or bowing of the column.

• Lagging: Inspect timber lagging for decay and steel lagging for corrosion or section loss.
Look for openings where backfill pressure is pushing material through joints or seams.
Inspect for signs of ground or rock failures behind the lagging.
• Lining: For concrete linings, inspect for new cracks, spalling concrete, and bulges or
displacement within or among adjacent panels. Inspect for water seeping through the
lining due to ground water pressure. Inspect the weep holes to ensure they are not
plugged, or have become the home for animals. Inspect for spalled pieces of lining that
may have fallen on the tunnel floor, or for fresh breaks along the top of the lining
indicating that pieces have spalled and fallen off. For masonry liners, inspect for cracks,
spalling, delamination of the outer exposed units, and for bulges or displacements,
settlement, or excessive mud pumping and fouled ballast. See Figures 11-17 & 11-22.

• Tunnel floor: Inspect the track surface through the tunnel shaft for indications of ground
heave.

• Inspect the tunnel walls, portals, sets, lagging or lining for signs of train impact caused by
clearance problems. If clearance monuments are in place along the tunnel, the inspector
must verify dimensions to the top of rail and center of track from the monument. See
Figure 11-23.

TUNNEL INSPECTION CHECKLIST

Portals
• Soil/rock stability
• Condition of headwall and wingwalls
• Drainage behind headwall and wingwalls
• Surface drainage

Shaft
• Clearances – Outcroppings or bulges in ceiling or walls
• Fractured or scaling stone in exposed rock tunnels
• Condition of rock bolts and anchor plates
• Condition and uniformity of sets and lagging on timber and steel linings
• Condition of concrete or masonry lining
o Bulges
o Cracks
o Spalls
o Tuckpointing
o Surface voids
• Tunnel floor
o Debris on floor indicating problems above
o Buckling or heave
o Fouled ballast
o Drainage
o Water/ice accumulation
Emergency

If the bridge inspector encounters a tunnel condition that affects the safe
operations of trains in, or near, the tunnel, contact the railroad dispatcher and
responsible authority to stop any trains and arrange for immediate repairs.
Criteria for the limits of acceptable/tolerable conditions must be provided by
the Railroad Owner or their designated engineer based on the principles of
acceptable rating contained in the appropriate chapter of the AREMA
Manual for Railway Engineering.
Tunnel Inspection

Figure 11-1 Fractured Rock above Portal Figure 11-2 Portal Damaged by Rock Falls

Figure 11-3 Poor Drainage and Ice Accumulation Figure 11-4 Tunnel Drainage Repair

Figure 11-5 Tunnel Portal with Headwall and Figure 11-6 Unstable Rock at Portal
Wingwalls
Figure 11-7 Common Headwall Layout Figure 11-8 Headwall Supporting Soil

Figure 11-9 Crack and Displacement of Headwall Figure 11-10 Drains in Snowshed Wall Adjacent to
Tunnel

Figure 11-11 Timber Shed at Tunnel Portal Figure 11-12 Rock Tunnel
Figure 11-13 Anchor Bolts Figure 11-14 Cribbing Material Behind Failed Liner

Figure 11-15 Concrete Lined Tunnel Figure 11-16 Brick Lined Tunnel

Figure 11-17 Failed Steel Liner Plates Figure 11-18 Failed Material Collecting on Tunnel Floor
Figure 11-19 Loose Rock Bolts Figure 11-20 Steel Sets and Lagging

Figure 11-21 Steel Lagging Between Sets Figure 11-22 Failed Concrete Liner

Figure 11-23 Impact Damage at Portal


CHAPTER 12
CULVERT INSPECTIONS

General

A culvert is a transverse drain or waterway under a road or railroad; a conduit, typically under
ground. What structures are defined as culverts varies from railroad to railroad. Most often the
determination is based on span length and occasionally on the type of construction. What the
structures all have in common is they are usually smaller structures built to handle storm water
runoff, intermittent streams or small waterways.

There are three typical culvert configurations

• boxes
• circular structures
• arches

All three types are constructed from a variety of materials. The primary materials used for
culverts are:

• Cast Iron Pipe


• Concrete Pipe
• Corrugated Wall Steel Pipe
• Concrete
• Masonry, Cut Stone or Brick
• Timber
• Rail
• Clay Pipe

A culvert may consist of any one of the materials listed, or it may be a combination of materials
and configurations.

Safety

Before considering inspection procedures, the inspector must first consider their safety. Culverts
are most often located in hard to access areas. Some may be considered to be confined spaces
(See Chapter 2). The inspector must determine the safest access to each location. He must decide
if it is a confined space and be completely familiar with all the rules and regulations that impact
the work he is doing. His personal safety and the safety of those working with him must be his
first and foremost concern.

For each type of culvert and culvert material there will be unique problems and deterioration or
failure modes. In general, all culvert inspections are basically the same and use the same tools. A
good culvert inspection does not require any high tech equipment but relies on the inspector’s
skills as an observer. (Occasionally robotic equipment does have to be used in confined spaces.)
The most useful tool is a good flashlight. Since many culverts will be too small for a “walk
through” inspection, the inspector will rely heavily on what he can see with the flashlight. Even
in larger size culverts there is seldom sufficient light for a through inspection.

Inspection

The basic inspection of all types and configuration of culverts is the same. Look for external
clues of failures and how the culvert is functioning. Observe the embankment above the culvert
and look for holes in the embankment or slope failures. Inspect the culvert for internal signs of
failure or conditions that might cause failure.

A culvert inspection does not start with the culvert - it begins at track level. First, inspect the
track and embankment above the area of the culvert for:

• Track alignment
• Surface problems
• Holes in the track structure. See Figure 12-1.
• Loss of Ballast

Any of these conditions could indicate a problem with the culvert. See Figure 12-2.

Observe the drainage channel that is upstream from, through, and downstream from the culvert
for the following conditions:

• Ponded water at the inlet. See Figure 12-19.

• Upstream for conditions that would effect the culvert such as drift, scour, change in
channel alignment. See Figure 12-3.

• Whether the inlet is open and unobstructed. See Figure 12-4.

• Look through the culvert to see if the line is smooth and open without erratic changes in
the alignment. See Figure 12-5.

• Look at the embankment at the inlet for high water marks. Culverts are most efficient
when they operate with some headwater depth, but indications of headwater depth greater
than one pipe diameter may indicate the culvert is undersized for current flow rates. If
there are indications of large changes in flow, check upstream for changes in land use that
may bring more water to the culvert at a faster rate. Excessive headwater depth might
saturate the fill, resulting in an embankment failure.

• Erosion downstream from the culvert and/or undercutting of the outlet structure. See
Figures 12-6, 12-7 & 12-18.
Material Specific Inspection Requirements

As previously noted, with each type of culvert there are some specific and unique failure modes
to check for.

Jointed Pipe

Concrete pipe, cast iron pipe and clay pipe are generally made with bell and socket ends and cast
in short segments. As a result there are numerous joints along the lengths of these culverts and
each joint is a potential failure location. Settlement of the embankment could cause the joints to
open (mismatched joints), resulting in water getting behind or under the culvert and/or allowing
embankment material to filter into the culvert. Each of these could lead to loss of embankment
material and ultimately loss of track surface.

Cast iron pipe and clay pipe are brittle materials and often fracture under load. This causes
portions of the culvert to break out with similar results as open (mismatched) joints.

Concrete Pipe is easily abraded by hard granular sediment that might be carried in the water and
reacts with the chemicals in the water. Both of these expose the wire reinforcing and could cause
a failure in the culvert.

Metal Pipe Culverts

There are several types of metal pipe culverts in use:

• Corrugated Steel Pipe – This could be round or elliptical in shape. It could be annular
ring riveted construction, or spiral welded. It comes in a wide range of sizes and is
generally installed by “open cut” methods. See Figure 12-8.

• Multi-plate or Tunnel Liner – It is typically constructed from curved steel plates in larger
diameters than corrugated steel pipe. This could be manufactured as corrugated, as two
flange or four flange plates, and is often used when the installation is made using
“tunneling” methods.

• Smooth-wall Steel Pipe – This is welded steel pipe without corrugations or joints. It has
improved flow characteristics because of the smooth walls. It is generally installed by the
“jack and bore” method.

Metal culverts are not as prone to joint separations as “Jointed Pipe” but share some similar
failure mechanisms:

• Metal pipes can be and are often abraded away due to hard granular sediment carried in
the flow.

• Metal pipes can also be eroded away due to acid or chemicals in the water.
In corrugated pipe the section loss from abrasion or acid or chemicals usually occurs in the
valleys first. This condition may go unnoticed because the peaks look good and the valleys are
obscured by sediment, ponded water or the water flow. This failure mechanism is also true in
multi plate, tunnel liner, and smooth wall pipe. Steel pipes should be closely observed for loss of
coating/galvanization to determine if either type of deterioration is taking place. This condition,
if left uncorrected, can result in erosion of the material under the culvert, or failure of the culvert
in compression, causing the culvert to collapse.

Flexible pipes require compacted fill to maintain shape and strength. An “arching” action of the
soil above the pipes also contributes to their capacity. Metal pipes, particularly corrugated metal
pipes, can deform from the circular form to flattened ovals when there is insufficient cover or the
soil adjacent to the pipes is improperly compacted or otherwise unstable. If the deformation
becomes great enough, the roof can collapse, the joints open up, and severe subsidence of the
track above can occur. See Figure 12-9.

Box/Arch Culverts

Cast-in-place Concrete Box and Concrete Arch Culverts

Cast-in-place concrete box and concrete arch culverts are among the most dependable types of
culvert construction because:

• They can be constructed as a single line or used in multi-line installations.

• Due to construction techniques and practices they typically have good foundations and
are constructed with thick sidewalls, floors and tops.

Concrete culvert structures develop problems similar to other concrete construction exposed to
weather. Freeze/thaw cycles cause cracking. Chemicals in the water react with the concrete
causing deterioration of the concrete and exposure of the reinforcing. Chemical reactions within
the concrete may cause the concrete to easily erode, exposing the reinforcing. As the process
continues, it accelerates and could result in holes in the floor, sidewalls or top.

Not all boxes and arches contain floors. If no floor is part of the design, undermining of the
sidewall foundations from excessive or rapid flow could cause uneven settlement, cracking, and
voids possibly resulting in the loss of track surface. If possible, concrete boxes and arches should
be sounded with a hammer to determine the condition of the concrete. See Figures 12-10, 12-11
& 12-12.

Masonry Culverts

Masonry culverts were some of the earliest culverts used on the railroads. They make use of cut
stone or brick built in box or arch shapes. Some masonry culverts were constructed with floors,
but many were not.

In cut stone culverts there are two primary modes of failure:


• The sidewalls become undermined.
• The capstones break. See Figures 12-13 & 12-14.

The two types of failure are generally found together. Undermining of the sidewalls causes
settlement. This settlement causes changes in the roof loading, resulting in capstone failures. The
capstones fall into the culvert and can reduce or block the flow. The hole in the roof allows the
embankment to fall into the culvert, resulting in track settlement.

The failure mechanisms of brick culverts are very similar to those of masonry culverts. Most
brick culverts are arch designs. Undermining of the foundation will result in changes in loading
causing the bricks in the arch to fall out, with the same results as those described for cut stone.

Inspection of masonry culverts should include a good examination of the foundations for
undermining and the mortar joints for full tight joints. Eroded mortar joints may cause an
otherwise adequate culvert to fail by allowing the masonry units to shift and/or drop out.

Timber and Rail-topped Culverts

Timber box culverts and rail top box culverts are very similar. Each typically has masonry
sidewalls and no floors. Timber boxes have large dimension lumber for tops, where rail top
culverts use interlaced rail to create the roof. Failure in these culverts is generally related to
undermining of the sidewalls or decay of the top material.

Attachments

Many culverts are built with headwalls, wingwalls, and aprons made of concrete, masonry, or
timber. Failure of these attachments can result in blockage of the culvert, impounding of water,
and/or causing an embankment failure. Often wingwalls and aprons break off because of being
undermined. Wingwalls frequently fail as a result of changes in the track embankment, such as
multiple track raises over many years, putting extra pressure on the wingwalls. See Figures 12-
15, 12-16 & 12-17.

Conclusion

Culvert inspections have many similarities regardless of material or configuration. Track


structure, embankment structure, culvert structure, stream characteristics and land use all play an
important role in the health of a culvert. Since the failure of a culvert can result in loss of the
roadbed, property damage or even a serious derailment, inspectors must be certain to look at all
the factors.

CULVERT INSPECTION CHECKLIST

Safety
• Proper authority and notification as needed
• Necessary tools to make the inspection
• Is this a confined space?

Track
• Alignment
• Surface
• Ballast section

Embankment
• Erosion
• Holes in the embankment
• High water marks

Stream condition
• Drift
• Scour
• Ponded water
• Channel changes
• Changes in land use

Culvert conditions
• Inlet/outlet – obstructed or blocked
• Scour
• Culvert – open or blocked
• Alignment
• Joint conditions
o Leaks
o Separated sections
o Most serious under track(s)
• Sidewall conditions
• Top/bottom conditions
• Missing/shifted stones or bricks
• Exposed reinforcing or deteriorated concrete
• Steel corrosion

EMERGENCY

If the inspector discovers a culvert condition that affects the integrity of the
culvert under train loads, or which can result in embankment or track failure,
contact the railroad dispatcher and responsible authority to stop any trains
and arrange for immediate repairs. Criteria for the limits of
acceptable/tolerable conditions must be provided by the Railroad Owner or
their designated engineer based on the principles of acceptable rating
contained in the appropriate chapter of the AREMA Manual for Railway
Engineering.
Culvert Inspections

Figure 12-1 Inspecting Track Above Culvert Figure 12-2 Embankment Failure at Culvert

Figure 12-3 Check Upstream Drainage Area Figure 12-4 Obstructed Inlet

Figure 12-5 Check Alignment Inside Culvert Figure 12-6 Undermining of Outlet Apron
Figure 12-7 Undermining of Outlet Apron Figure 12-8 Corrugated Steel Pipe

Figure 12-9 Deformation of Metal Pipe Figure 12-10 Undermining of Box Culvert Floor and
Sidewall

Figure 12-11 Undermining and Failure of Culvert Figure 12-12 Steel Pipe Placed Inside Failing Box
Floor Culvert
Figure 12-13 Undermined Masonry Sidewalls Figure 12-14 Heavily Damaged Masonry Wingwalls

Figure 12-15 Headwall and Wingwall Displacement Figure 12-16 Separation of Headwall from Barrel
Due to Slope Failure

Figure 12-17 Structural Failure of Headwall Figure 12-18 Drop Inlet to Culvert
Figure 12-19 Erosion at Culvert Inlet
CHAPTER 13
EMERGENCY BRIDGE INSPECTIONS

General

Many events can trigger the need for an emergency bridge inspection. The most common
include: fire, flood, derailment, impacts by motor vehicles or vessels on waterways, and
catastrophic events such as hurricanes or earthquakes.

Safety is a primary concern for any inspection. Prior to accessing or inspecting a bridge
following an emergency, make sure the site is safe. Check with authorities for the
possible presence or release of any flammable or hazardous materials. Be on the lookout
as you move around the site and bridge for unstable structure or debris (underfoot and
overhead), sharp objects, snakes or other vermin, and other potential concerns based on
the type of emergency. Make sure all appropriate personal protective equipment is being
used.

An emergency inspection has several purposes. Initially, it must be established if the


structure is capable of safely supporting train traffic. A careful evaluation of all the
critical components of the bridge must be made prior to making this determination. In
some cases it may be decided that train traffic can resume, but at restricted speed until
repairs are made. The inspector needs to identify immediate repairs to be made prior to
resumption of train traffic, and additional repairs required to restore the structure to its
original or desired condition.

Some repairs and requirements to make a bridge safe for train operations will be obvious.
Other damage may require considerable expertise and/or experience to evaluate. It is
very important that the inspector know his/her limitations and request further help in the
evaluation from a qualified supervisor or engineer when necessary.

As with virtually all bridge inspections, a good place to start is to observe the line and
surface of the track. Any deviations from normal indicate a potential problem and
identify a good location to begin inspecting for damage.

Fire

When evaluating the damage from a fire, it can be helpful to get any available
information about the intensity of the fire. Eyewitnesses may be able to provide valuable
information. Some materials on site may have melted and can give a clue about the
temperature such as: lead 620ºF, plastics 300-450ºF, glass 750-900ºF, aluminum 1200ºF,
and copper (rail bonding wire) 2000ºF. Check to see if the rails are kinked or warped due
to the heat.

• Timber Fire damage is easily evaluated on timber structures, but that evaluation
can be quite time consuming. The damage caused by fire is often not quite as
severe as the actual event would seem to indicate. The only way to ascertain the
true extent of damage is to chip away the charred material at a number of
locations and measure the section of good lumber that remains.

Locations where two or more members come together often experience the
greatest section loss. Look closely at pile to bracing connections and cap to
stringer connections. Stringers with spacing washers between them often exhibit
significant section loss due to a “chimney” effect between the stringers during a
fire. Critical locations to check on stringers are at the bent caps and mid-span.
Ties on timber trestles should be evaluated to see if there is sufficient section left
below and around the tie plates to transfer the load to the stringers. Ties on steel
stringers or girders need to be checked for remaining cross section and evaluated
for shear and bending capacity.

• Concrete Concrete structures exposed to fire generally perform quite well but,
depending on the intensity and duration of the fire, permanent damage can occur
at temperatures above 570ºF. As temperatures go above 570ºF concrete will have
discoloration progressing from pink to white to a grey-buff at very high
temperatures. Note any discoloration.

Sound areas of concrete that have been exposed to fire with a hammer to check
for any delamination or damage. Check for cracking, distortion, spalling or any
other indications that damage may have occurred. Note any exposed reinforcing,
particularly in prestressed or post-tensioned members. See Chapter 8, Part 21 of
the AREMA Manual for Railway Engineering for more information related to fire
damaged concrete.

• Steel Steel that has been damaged in a fire should be carefully examined for
evidence of deformation. Check the straightness of beams, columns, and other
members. On plate girders, the girder will often look straight but there may be
localized deformation of the web plate between stiffeners. Note the amount and
location of any deformations. Members exposed to extreme heat that would
permanently reduce their strength capacity are typically so badly deformed that
they will require replacement anyway.

Connections should be carefully examined for damage caused by thermal stresses


created by the heating and cooling of steel members. Look for sheared or loose
rivets or bolts as well as cracked welds. It may be necessary to obtain coupons of
the steel from appropriate locations for laboratory testing of mechanical
properties. See Chapter 15, Part 8 of the AREMA Manual for Railway
Engineering for more information related to fire damaged steel.

Flood

The most common concern from flooding is the undermining of the substructure. It is
helpful to know during the inspection whether the foundation is on spread footings, or a
deep foundation such as piles or drilled shafts which are considerably less susceptible to
scour. If the structure is accessible during the flood event, watch for areas with drift
accumulation or where hard currents or eddys are occurring, these are likely locations for
scour. Check the line and surface of both the track and structure, taking particular note of
areas with scour potential.

It is not uncommon on bridges with tall bents and significant drift accumulation for the
bridge to be moved out of alignment during high flow. The bridge may move back into
line after the flow subsides, but the inspector should look closely for signs of damage in
connectors and slender members such as piles.

Note the condition of the embankment immediately adjacent to the structure abutment
and backwall.

Check for localized impact damage from drift, other floating debris, or even vessel
strikes.

Derailments

First determine if there are any derailed cars of hazardous materials on the train. If so, do
not approach the site until the area has been declared safe by an authorized individual.
Use extreme caution to determine if the bridge and derailed equipment is stable. All or
some of the inspection may need to be delayed until clean-up operations are underway or
completed. If the inspection is performed during clean-up, do a thorough job briefing
with the other workers involved and keep yourself safe. Be aware that additional damage
is often caused as part of the clean-up operation.

Carefully inspect the structure for damage. The damage may be caused by direct impact
from derailed equipment or by the unusual and random impacts and loading that occurred
during the derailment. Look closely at bearings, connections, and other locations subject
to prying or pulling apart. Also, on steel bridges, carefully inspect fatigue sensitive
members such as truss hangers. Damage may not be limited to the span(s) where the
derailed equipment rests. Look over other portions of the structure that could have been
affected by the movement or shifting of spans.

Impact Damage

Impacts from motor vehicles or vessels on waterways are not uncommon. The first step
in the inspection is to check and see if the structure has been knocked out of line.
Misalignment can be caused by one or more of several factors:

• Tie deck fasteners broken and track shifted


• Deformation of the superstructure or span caused by the impact
• Span shifted on bearings or bearings shifted on bridge seat
• Movement of the substructure
Determine the cause of the misalignment and whether substantial repairs are required, or
just realignment. Identify any specific structural damage to the superstructure, bearings,
or substructure.

On movable bridges the damage may also include pier protection cells or fendering,
dolphins, shear fencing, machinery, and navigation lights. Even if, following an impact,
machinery damage is not apparent, the bridge machinery should still be operated through
its typical operational cycle to be sure all machinery, locks, limit switches, signal
interlocks, etc. are operating properly.

Catastrophic Events

Certain events such as floods, earthquakes, and hurricanes may involve many structures.
In these cases, the scope and breadth of the inspection should be based on standard
policies or procedures that are established prior to the event, research to determine the
area of potential damage, and an inspection methodology or procedure based on the
anticipated type of damage. A separate chapter on Post-Earthquake Inspection is
provided elsewhere in this Handbook.

EMERGENCY BRIDGE INSPECTION CHECKLIST

Site Safety & Access


• Check for hazmat
• Fire or hot spots
• Flooding
• Snakes and vermin
• Structure stability

Track
• Track alignment and surface profile. Determine if any defects are track or
structure related.
• Condition of rail, fasteners, and other track material (OTM)
• Condition of track or bridge ties
• Ballast
o Adequate shoulder
o Ballast retainers
o Locations that are missing ballast

Structure Damage
• Repairs required to restore traffic
• Long term repairs required
• Can trains run
• Can trains run at reduced speed
• Will cleanup cause additional damage
Repairs
• Who will make repairs
• Material and equipment needed for repairs
• When can train traffic resume
CHAPTER 14
POST-EARTHQUAKE INSPECTIONS

General

Chapter 9 of the AREMA Manual for Railway Engineering presents items and issues that should
be considered by a railroad when developing its procedure for post-earthquake inspections.
Inspections should follow the railroad's own procedure when provided, but the following list is a
general guideline that may be used if a more detailed procedure is not otherwise available. Items
of concern which are not directly related to structural inspection are included, because structure
inspectors might well encounter these items during a post-earthquake inspection.

Track and Roadbed

• Line, surface and cross level irregularities caused by embankment slides or


liquefaction, track buckling or pull aparts due to soil movement, offset across fault
rupture, etc.
• Disturbed ballast.
• Cracks or slope failures in embankments.
• Slides and/or potential slides in cuts, including loose rocks that could fall in an
aftershock.
• Scour due to tsunami in coastal areas.
• Potential for scour or ponding against embankment due to changes in water courses.

Drainage

• Blockage of cut ditches or other changes in drainage patterns. (While these conditions
will not usually prevent restoration of service, they will require correction.).

Steel Bridges

• Displaced or damaged bearings.


• Stretched or broken anchor bolts.
• Distress in viaduct towers.
• Buckled columns or bracing
• Tension distress in main members or bracing.
• Displaced substructure elements.

Concrete Bridges and Substructures

• Displacement at bearings.
• Displaced substructure elements.
• Cracks in superstructure.
• Cracks in substructure.
Timber Trestles

• Line, surface and cross level of track. (Movements that do not affect line, surface or
cross level are unlikely to be damaging, especially in open deck trestles).
• Displaced timbers, particularly in framed bents.
• Broken bracing.
• Bent bolts or drift bolts.

Movable Spans

• Damage to counterweight guides.


• Open swing span shifted on pivot pier.
• Relative movement of piers that prevents opening or closing, including misalignment
of track girders and segmental girders of rolling lift spans.
• Machinery damage.

Culverts

• Damage to culverts is unlikely if line and surface of track are good and no slides or
embankment spreading are observed.

Retaining Walls

• Increased tilt in walls which may be caused by footing rotation or cracking at base of
stem. (Walls with surcharge above top of wall appear particularly vulnerable.).
• Line deviations due to sliding.

Tunnels

• Fallen material or loose material that may fall in an aftershock.


• New cracks or failures in lining.
• Offsets due to displacement across fault.
• Unusual flow of water within tunnel.

Other Structures

• Structural and/or non-structural damage to essential buildings that would prevent or


inhibit use. NOTE: Inspect promptly, with concurrence of local building authorities,
to prevent outside inspectors from “red tagging” buildings that are damaged but not
unsafe.
• Leaks and/or structural damage to fueling facilities, including tanks and pipelines.
Look for evidence of leaks in buried fuel lines.
• Catenary support structures and tension-regulating systems of electrified lines.
NOTE: Electric traction power substations should be inspected by a qualified
individual.
Structures That Could Fall On the Track

Overpasses

• Reduced support for span at bearings.


• Column damage.
• Damage to any span restraint system.

Adjacent Buildings

• Structural damage affecting ability to resist aftershocks.


• Clearance infringements.
• Power lines that may be vulnerable to aftershocks.

Signal and Communications Facilities

Signal and communications facilities must be inspected by qualified personnel. However, others
involved in inspection should note damage to pole lines and other obvious damage to equipment.
Signal masts, signal bridges or instrument housings observed to be out-of-plumb should be
reported immediately.
GLOSSARY
Abutment - That part of a pier from which an arch springs. A structure sustaining one end of a
bridge span and at the same time supporting the embankment which carries the track or roadway.

Aggregate - The inert material such as sand, broken stone, etc., with which the cement or other
adhesive material is mixed to form a concrete or mortar.

Anchor Pier - A pier used in cantilever bridges to resist the uplift at the end of the anchor arm.

Arch - Any bow-like curve, structure, or object, usually having the convex side upward, generally
spanning an opening and producing horizontal as well as vertical reactions.

Arch Bridge - A curved structure which produces reactions inclined to the vertical.

Arch Culvert - A culvert having an arch roof.

Axle Load - The load which comes on an axle of a car, or locomotive and is in turn transferred to
the structure through two wheels.

Babbitt Metal - An alloy of tin with copper and antimony, used for lining bearings and making
bushings.

Back Wall (or Head Wall) - The wall above the bridge seats on abutments at the end of a
structure.

Ballast Deck - A bridge floor under a railway track upon which ballast is placed with ties
embedded therein.

Ballast Retainer - A timber, concrete, or steel riser on both edges or ends of a ballast deck that
keeps ballast from falling off of the bridge.

Bank Protection - The prevention of erosion of a bank of a stream by the use of riprap, mattresses,
or other artificial means.

Bascule Bridge - A bridge having a span that opens by rotating in a vertical plane.

Base Casting - A steel or iron casting upon which the bridge-shoe rests.

Base of Rail - The bottom of any rail placed in final position. It generally determines the elevation
from which the heights of the various parts of the structure are measured.

Base Plate - The foundation plate of metal on which a heavy piece of machinery or the end of a
bridge rests. This plate is usually set on masonry or concrete. Also called a masonry plate.

Batten Plate - A stay plate at the ends of a compression member. Sometimes termed tie plate.
Batter Pile – The outside pile of a bent that is driven at an angle from vertical to provide lateral
stability to the bent.

Beam - A member the principal function of which is to carry a transverse load, or more simply, a
member stressed primarily in bending.

Bearing Pile - A pile which obtains its primary load capacity through end bearing on a hard
stratum of soil or rock at the pile tip.

Bearing Pin - A truss pin at the end of a span connecting the truss to the shoe.

Bending Moment - The moment which produces or tends to produce bending in a beam or other
member of a structure. It is measured by the algebraic sum of the products of all the forces by their
respective lever arms.

Bent - A supporting frame consisting of posts or piles with bracing, caps, and sills.

Blocking - A set of timber blocks which is placed under bridge bearings or members to raise and
support them.

Boat Spike - A square, chisel-pointed spike with a rounded head, ordinarily from eight to ten
inches long, used to fasten heavy planks in wooden floors, railroad crossings, etc.

Bore - The internal diameter, of a hole, tube, or pipe.

Bottom Lateral Bracing - Lateral bracing in the plane of the bottom chords of truss spans or
bottom flange of girder spans.

Box Beam or Girder - A hollow beam, generally rectangular in section, having its sides made of
plates united by angle-irons or welds, or a concrete beam with a rectangular cross-section and a
hollow core.

Box Culvert - A square or rectangular shaped culvert.

Brace - Generally a strut supporting or fixing in position another member.

Brick Masonry - Masonry composed of brick.

Bridge Frog - A contrivance built of two or more pieces of rails mounted on a common base and
used for passing the car or locomotive wheels across the ends of a movable bridge.

Bridge Seat - That part of the top of a bridge pier or abutment where the pedestals or bearings of
the superstructure rest.
Built-up Beam or Girder - A beam or girder made up of structural shapes, such as plates and
angles, riveted, bolted or welded together.

Buttress - A short cross-wall built against the main wall to increase its stability.

Caisson - A bottomless box or enclosure, surmounted by a crib or shaft which is excavated at the
bottom to sink the entire structure to a required depth to form the foundation for a pier structure.

Camber - The upward curvature of a span above its nominal position.

Cantilever Bridge - A structure at least one portion of which acts as an anchorage for sustaining
another portion which projects beyond the supporting pier.

Cap Plate - The top plate on a steel column or post.

Cast Steel - Steel that is cast into shape directly from the furnace instead of being cast into ingots
and rolled or melted.

Cement - Hydraulic calcium silicates that react chemically with water and are combined with
aggregate to form concrete.

Center Bearing Swing Span - A term applied to swing spans to indicate that the dead load support
is near the axis of the pivot pier instead of near the outer edge.

Centerline – Centerline of the track, used as the horizontal reference

Centrifugal Force - The outward force exerted by a train going around a curve due to its inertia,
against that force which is causing it to deviate from a straight-line motion and to travel in a curved
path.

Chamfer - To bevel a sharp edge or corner, typically on concrete.

Channel - (1) The deepest Part of a river or stream; usually that part available for navigation. (2) A
structural or rolled steel shape used in bridge building and in other steel construction.

Check - A small crack in wood due to seasoning, or in concrete or mortar due to drying.

Chord - (1) The primary members on the top or bottom of a truss span that run the full length of
the span parallel to the track(s). The top chord is usually in compression and the bottom chord is
usually in tension (except on continuous spans). (2) Timber stringers bolted together to form one
larger timber beam.

Chord Bolt - A bolt through individual timber stringers fastening them together to form a chord.

Clearance Line - A line on a diagram showing the minimum clearance allowed.


Cofferdam - A temporary enclosing structure, practically watertight, from which the water is
pumped to create a safe working area.

Collision Strut (post) - A short, diagonal strut used to provide support to a truss end post.

Column - A vertical compression member which supports a part of a bridge.

Composite Steel Bridge - A steel span with a concrete deck rigidly attached to the steel using
shear connectors so both the steel and concrete carry the bending and, in some cases, shear forces.

Compression - The state of being compressed; shortening by pressure.

Concrete - A material composed of cement, sand, gravel, and water.

Continuous Span - A span that rests on three or more supports with main members being
continuous over one or more internal supports.

Cope - To notch steel beams, channels, etc.

Coping - The top or cover of a wall, column, or pier. Usually made so as to project beyond the
face below.

Corbel - A small shelf cantilevered out from a beam, wall, or column in order to support a beam or
a superincumbent load.

Corbel Block – Short timber blocks placed on pier caps or subcaps to provide support for
additional caps or bridge bearings.

Corrosion - The disintegration of a substance by the action of chemical agents.

Counter - One of a pair of diagonals placed in a truss panel, in the form of an X, where a single
diagonal would be subjected to stress reversals.

Counterweight - A weight that counterbalances span weight on a movable bridge, used on lift
spans, bascules, swing spans, etc.

Course - A horizontal layer of stone in a masonry wall or substructure unit such as a pier or
abutment.

Cover Plate - A plate fastened on the flanges of a girder to give additional cross-section thereto; a
top or bottom plate of a chord member.

Creosoted Timber - Wood which has been preserved through a pressure-treatment process using
creosote, where penetration of wood with preservative provides long-term protection against decay.

Cribbing - Timbers piled cross-wise in order to form a support for a load.


Cross Frame - A transverse bracing frame between stringers or girders. Also termed a "Buck
Brace."

Cross Girder - Any girder passing across a bridge from one truss or main girder to another, and,
generally, perpendicular to the truss or girder planes.

Cross Level – The vertical position of one rail respective to the other. On tangent track both rails
should be at the same elevation.

Cross-over - An arrangement of turnouts enabling movement from one track to another

Crossing - An intersection. The place where two railroads cross or a roadway crosses the tracks.
The term is also used for a bridge crossing a stream, river, railroad, or highway.

Cutoff - That part of a pile that has been sawed off after the pile is in place.

Cutting Edge - An edge of timber or steel angles placed on the bottom of the working chamber of
a caisson.

Cutwater - A starling; the projecting ends of a bridge pier, etc. Usually so shaped as to allow
water, ice, drift, etc. to strike without injury to the structure.

Cylinder Pier - A pier made of a cylindrical steel shell filled with concrete.

Dead Load - The weight of all the parts of a bridge itself and anything that may remain upon it for
any length of time, such as tracks, walkways, utilities, etc.

Deck - The flooring of a bridge.

Deck Bridge - A bridge where most or all of the structure is below track level.

Deck Plate Girder - A deck bridge fabricated from steel plates and angles riveted, bolted, or
welded together.

Deck Truss - A truss span where the entire structure is at or below track level.

Deflection - The vertical displacement of the bridge or a bridge member caused by loading.

Deformation - A change of shape in a member due to applied loads.

Diagonal - A member running at an angle across the vertical panel of a truss.

Diagonal Bracing - Bracing along diagonal lines.


Diaphragm - A stiffening plate or section used in the interior of a column or between bending
members to give them additional strength and rigidity.

Dolphin - A cluster of piles driven some distance ahead of the ends of the channel span piers of a
bridge to protect the faces of the piers against blows from passing vessels.

Double Cap - Two caps set on top of one another.

Draw - The movable portion of a draw-bridge.

Draw Bridge - A movable bridge that may be drawn or turned to one side, or lifted up, either
bodily or in sections, so as to permit boats to pass under or through it.

Drift Bolt - A short rod or square bar to drive into holes bored in timber for attaching adjacent
members to each other or to piles. The length generally varies from one foot to two feet. A drift
bolt is generally provided with some sort of head. Drift bolts with a sharpened end are often
referred to as drift pins.

Drip - A small channel cut under the lower projecting edge of a coping, etc., so that when rain
reaches that point, it will drip or fall off.

Drum - A revolving cylinder around which ropes or belts either travel or are wound, such as on a
movable bridge.

Drum Girder - The circular, main support girder at the center portion of a swing span.

Dump - The backfill area immediately behind a bridge abutment.

Dump Bent - The end bent on a timber bridge.

Earth Pressure - The lateral pressure exerted by a bank of earth when supported by a retaining
wall or an abutment.

Eccentric Load - A load which is applied off-center to the axis of a member, producing a bending
moment on the member considered.

Eddy - A whirl or backward current of water. A vortex. That portion of the water in a stream that
actually swirls.

Efflorescence - A powder-like incrustation formed on the surface of concrete.

Elongation - The stretching or extension of a part beyond its natural dimensions.

Embankment - A bank, a dike, or an earthwork raised for any purpose.

End Floorbeam - The floorbeam at the end of a span.


End-lift Machinery - The machinery that releases the ends of a swing span for turning.

End Post - The post at the end of a truss. Also referred to as the batter post.

End Stiffener - Vertical angles fastened to the web of a plate girder at its ends for the purpose of
stiffening it and transferring the end shear to the shoe or base plate.

Erosion – The lateral loss of stream bank material

Expansion Bearing - A support at the end of a span where provision is made for expansion and
contraction of the structure. An expansion bearing may also allow rotation to accommodate span
deflection.

Expansion Joint - A joint in which movement for expansion and contraction is allowed.

Expansion Rollers - A group of steel cylinders nested in a box or suitable frame placed under the
shoe of a span to facilitate its movement during temperature changes and loading. (See roller nest)

Eye-bar - A bar with an eye at either one end or each end (also I-bar).

Fatigue Cycle - The loading and unloading of a bridge member during the passage of trains, and
having a sufficient range of stress to accumulate fatigue damage.

Fender - A guard for protection. Timbers, piles, etc., to protect vessels from striking, rubbing, and
scarring piers.

Fender Pile - A pile which is driven at wharfs, or in front of large masonry piers or other important
works, to protect them from sudden blows by vessels.

Fill Plate - A plate used to fill open spaces between members or components of members.

Fillet - A plain, narrow, flat molding in a cornice or a corner. The rounding of sharp corner.

Fish Plates – Splice plates used to join to beams on the webs.

Fish Bellied Girder - A girder having the top flange horizontal and the bottom flange curved to
rpovide a smaller depth at the end.

Fixed Bearing - A support at the end of a span where it is firmly connected to prevent any
longitudinal movement of the structure. A fixed bearing may allow rotation to accommodate span
deflection.

Fixed Bridge - A bridge in a fixed location that does not move except for expansion and
contraction.
Flange - Angles or plates at the top and bottom of a beam or girder which resist tension or
compression caused by bending.

Flange Angle - One of the upper or lower chord angles in a beam or girder; or the angles which
either alone or with cover plates make up the flange of a built-up beam or girder.

Flange Coupling - A coupling made up of two parts, each firmly attached to the end of its shaft,
bolted together to form a permanent connection.

Flange Splice - A splice made in the flange of a beam or girder.

Flexure - Bending.

Floorbeam - A transverse beam or girder placed at the panel points of a span to support the
stringers which carry the deck. In some instances on through plate girders, floor beams are closely
spaced to support the deck without stringers.

Floor System - The system of members in a bridge that carries the deck and its load, transferring
the loads to the main girders or trusses.

Footing - The spreading course or enlarged portion at the base of a pier or abutment.

Force - That which moves or tends to move matter. The action between two bodies either causing
or tending to cause change in their relative rest or motion.

Foundation - That portion of a structure, usually below the surface of the ground, which distributes
the pressure upon its support. Also applied to the supporting material itself.

Fracture - To break or split. A partial or total separation of parts of a continuous solid body under
the action of force.

Framed Bent - A bent consisting of a sill, posts, and a cap in contrast to a bent that is a cap on
driven piles.

Friction Pile - A pile in which the bearing capacity is mostly developed from the friction of the soil
surrounding it.

Gage - The distance between the inside faces of both rail heads; 4’8½” for US Standard Gage.

Gear - A wheel having teeth on its periphery or face. A piece of mechanism for transmitting
motion.

Girder - A primary beam or built-up member carrying loads to the bridge bearings or supports.

Girder Bridge - A bridge composed of plate or lattice girders.


Girt - Longitudinal brace on a timber bridge.

Grade - The degree of inclination from the horizontal, expressed usually in percentage.

Grade Crossing - A crossing where road and track are at the same elevation.

Grout - A mortar composed of sand, cement, and water of such liquid consistency that it can easily
be poured.

Guard Timber - A guard-rail made of a timber, bolted to or dapped over the ties for railway
bridges. Also referred to as spacer timber or tie spacer.

Gusset Plate - A large connecting plate used at panel points to join the chord and the web members
of a truss or bracing mmebers.

Hammer Head Pier - A pier consisting of a relatively slender shaft flaring to a wide top to
accommodate the bridge bearings.

Hangers - A hip-vertical or suspender of a truss acting in tension. Also a tension member


supporting a floor system in an arch or in a suspension bridge.

Hardpan - A very compact layer or bed of material under the track.

Heartwood - The oldest, central rings of any timber. Typically, preservative treatment will not
penetrate into the heartwood, making it susceptible to decay.

Hogjaw - A diagonal bracing strut from the bottom of one bent on a timber bridge to the top of an
adjacent bent to provide longitudinal stability.

Hydration - The process of combining or impregnating with water, or the resulting condition.

I-Beam - A rolled structural shape having a cross-section resembling the letter "I."

I-Beam Bridge - A small bridge consisting of a floor supported on I-beams.

Ice Guard - A fender placed at the up-stream end of a bridge pier to divert the ice or else to break
up the large floes into small pieces.

Ice-breaker - A structure of masonry or timber (as a pier or a cluster of piles) for the protection of
bridge piers against moving ice.

Impact Load – (1) A dynamic increment of load created by moving loads traversing a bridge. (2)
A short duration, often high magnitude load striking a portion of a structure. This can include flat
train wheels, wheels moving over rail joints, vessel or vehicle strikes, etc.

Inner Guard-rails - Guard-rails placed between the running rails of a track.


Interlocking - Signal appliances that are interconnected so that each of their movements follows
the other in a proper sequence.

Intermediate Floorbeam - Any floor-beam between the end floor beams.

Intermediate Sill - A horizontal member in the plane of a timber trestle bent between the
elevations of cap and sill, to which the posts are framed.

Intermediate Stiffener - Any one of the stiffeners on a plate girder between the end stiffeners.

Invert - The flow line of a sewer, culvert or tunnel.

Jacket - A layer of concrete placed over an existing pier or abutment surface to strengthen,
stabilize, confine, or protect it.

Jetty - A structure of wood, stone, or other materials extending into a body of water and serving for
a wharf or pier, or as a mole, rampart, or wall. Also used to restrain, charge, or direct a current, and
to protect a harbor, shore, channel or the like.

Journal - That part of a shaft or axle which rests on the bearings.

Key-way - A slot cut in a shaft or hub of a gear or pulley to receive the key.

Knee Brace - A short diagonal brace, used to connect or stabilize members against buckling and
out-of-plane bending.

Knife-edge - A sharp edge on corroded steel similar to that of a knife blade.

Lacing Bars - A system of bars not intersecting each other at the middle, used to connect two
members of a strut in order to make them act as one member.

Lag Screw - A large-sized wood screw with a square head larger than the shank for convenient
turning with a wrench, and having a special thread to increase the holding strength.

Lateral Bracing - A system of tension or compression members, or both, forming the web of a
horizontal truss connecting the corresponding chords or flanges of the opposite trusses or girders of
a span.

Lattice Bars - A system of bars crossing each other at mid-length, used to connect the two
members of a strut in order to make them act as one member. Generally the crossed bars are riveted
together at their intersection.

Lattice Truss Bridge - A bridge having riveted trusses with multiple intersection web systems.

Leaves - The portions of a moving bridge which revolve.


Lift Bridge - A style of movable bridge which travels in a vertical plane, sometimes called a hoist
bridge.

Line - The lateral, or side to side, tolerance of a section of track to its original survey. Also called
Alignment.

Live Load - A moving load on a structure, such as a train.

Load - The weight carried by a beam, girder, truss, span, or structure of any sort, including its own
weight.

Longitudinal Shear - A shear parallel to the longitudinal axis of a member.

Masonry - A general term applied to structures made of stone, brick, or concrete.

Masonry Joint - A joint between masonry stones that is filled with mortar.

Masonry Plate - A plate used under a bridge-shoe for the purpose of distributing the load on the
masonry.

Mattress - A combination of willow poles and wire rope woven together, forming a mat which is
placed on the bed or the bank of a stream to prevent scouring.

Mortar - A sand, cement, and water mixture used to fill the voids and transfer loads between
stones or bricks in masonry structures.

Mud Line - The soil/water interface in a profile of a river crossing.

Mud Sill - Timber blocking resting on the earth, to support a framed bent.

Nose - A pointed or tapering projection on the upstream or downstream edge of a pier, may act as
an ice-break.

Operator’s House - A bridge-tender’s house from which the operation of a movable span is
controlled.

Overhead Crossing - A crossing where a bridge carries a road or tracks over the railroad.

Packing Bolt - A bolt which holds together the several parts of a member, also called chord bolts.

Packing Diagram - The arrangement of eye-bars on a truss pin.

Panel - That portion of a truss between adjacent panel-points lying in the same chord.

Panel Length - The distance between two adjacent panel points in the same chord of a truss.
Panel-point - The point at which the axis of a principal web member intersects the axis of a chord
of a truss.

Parapet – (1) A raised wall or curb at the periphery of bridge spans or abutments to retain ballast.
(2) A wall-like step placed on bridge piers to accommodate different span depths.

Parting Line - The location over a bent or pier where two stringers butt up against each other end
to end on a timber bridge.

Pedestal - A footing that raises a bearing above the bridge seat, typically of steel or concrete.

Pier - A structure composed of masonry, concrete, steel, timber, or a combination of same which is
used to transmit the loads from a bridge superstructure to the foundation.

Pilaster - A thin, flat projection from the face of a wall made to resemble a column, for ornamental
purposes.

Pile - A long, heavy post or pole of timber, concrete, or steel driven into the ground to carry a
vertical load, resist a horizontal force or both. Some piles may be driven as hollow shells and later
filled with concrete or grout.

Pile Bent - A bent having piles for supporting posts.

Pile Cluster - Several piles driven close together forming a group or cluster.

Pile Foundation - A foundation formed in soft soil by driving a group of piles to a depth which
will give them the requisite capacity to carry the load.

Pile Pier - A pier formed by driving a cluster of piles and capping them in the form of a grillage to
carry the shoes of the span.

Pile Trestle - A trestle having pile bents for supporting the stringers.

Pin Plate - A plate riveted to the outside of the end of a member where it connects to a pin to give
additional strength and greater bearing on the pin.

Pin-connected Truss - A term applied to the method of joining the members of a truss by pins
instead of using riveted connections.

Pinion - Any toothed gear of small size as compared with the gear which it engages.

Pivot Pier - The pier supporting a swing span and upon which it turns.

Plan - The general layout of a structure.


Plank - A piece of lumber thicker than a board; usually measures from two to four inches in
thickness and from six inches upward in width.

Plate Girder - A girder built of structural plates and angles.

Plumb - Vertical.

Plumb Pile - A pile driven vertically, usually one of the inside piles of a bent.

Pony Truss - A low truss without any overhead bracing.

Portal Bracing - The combination of struts and ties in the plane of the end posts at a truss span
portal which helps to transfer the wind pressure from the upper lateral system to the pier or
abutment.

Post - A vertical, or nearly vertical, compression member.

Quarter Pile - A bent pile driven with some incline to the vertical, located between the interior
plumb piles and the exterior batter piles. Also called a rail pile, it is often located below the rail.

Rack-circle - A rack bent into the form of a circle that is engaged by drive pinions on swing or turn
spans.

Radial Strut - One of a series of struts radiating from a fixed point such as the radial braces of a
turntable, or a swing-span drum.

Rail-lift - A device used on swing spans for lifting the ends of the rails, so as to clear obstructions
on adjacent spans as draw is swung open.

Rail-lock - A device used on swing spans for locking the rails at the ends of the span after closing
the draw.

Reaction - A passive force set up in opposition to an initial, active force, e. g., the upward pressure
on the bottom of a beam resting on a support, equal in amount to the downward force from the
loads on a beam.

Reinforced Concrete - Concrete in which steel bars are inserted to strengthen it, principally by
resisting the tensile stresses induced by external forces.

Reinforcing Plate - An extra plate used to reinforce or strengthen a member.

Rest Pier - A pier which supports one of the ends of a draw span.

Retaining Wall - A wall built to sustain a lateral pressure, such as an earth thrust.

Revetment - A facing of wood, mattress, stone, or concrete placed to prevent erosion.


Rim Bearing Draw - A term applied to swing spans to indicate that the dead load is supported by a
circular girder near the periphery of the pivot pier instead of near its axis.

Riprap - A facing of stone, concrete, or planks placed on the bank slope of a stream or around a
pier to prevent erosion.

Riveted Truss - Any truss having its main members riveted together.

Rocker - A casting or built-up steel frame fastened to the end of a span or column to permit a slight
rotation.

Rocker Bearing - A bearing, or support, for spans which permits a slight rocking with the
changing position of the live load and with variations of temperature.

Roll Rack - A rack on which a pinion works.

Rolled Steel - Steel that has been cast into ingots and then passed through a succession of rolls
until the desired final shape is obtained.

Roller - Any short, round bar put under an object to facilitate its movement.

Roller Bearing - A shoe or plate resting on rollers which in turn rest on a base casting at the
expansion end of the span.

Roller Nest - A group of rollers, enclosed in a suitable frame or box, which support a bridge shoe.

Rolling Lift Bridge - A bascule bridge in which the moving arm rolls on a plane or upon friction
rollers.

Sap Wood - The outer and lighter colored portion of a timber.

Sash Brace - A horizontal member secured to the posts or piles of a bent between the cap and sill.

Scour -The general or local vertical deepening in normal stream bed elevation. Scour often occurs
around obstructions in a stream such as piers or abutments.

Seat Angle or Shelf Angle - A short angle fastened to a column or beam to temporarily support a
beam during erection.

Secondary Member - A subordinate part of a bridge. Generally refers to the suspenders and sub-
diagonals of trusses.

Shaft - A long, cylindrical bar capable of rotating and transmitting torque.

Shaft Coupling - Any of the several devices for joining the ends of two shafts.
Shakes - Splits or checks in timber which usually cause a separation of the wood between the
annular rings.

Shale - A hard, clay-like formation having a closely laminated structure.

Shear - To slide one part of a body upon an adjacent part. The stress set up in opposition to a
shearing action.

Sheave - A wheel with a grooved face for carrying a rope or cable.

Sheet Pile - A form of piling used to shut out water or retain earth, generally made of steel and
arranged to secure a tongued and grooved effect when driven close together.

Shim - A small piece of wood, metal, or other material placed between two parts or members of a
structure to bring them to a desired relative position.

Shoe - That part or detail of a bearing assembly which transfers the load from the end pin to the
bearing plate or to the intervening rollers. Also a point used on piles when driving them through
hard ground.

Shop Drawing - A drawing of a structure or machine showing all parts and dimensions so that the
shop can actually build what is indicated on the drawing without other information.

Silica - A dioxide of silicon (Si02). It occurs in nature as quartz.

Sill - The lower horizontal member of a framed bent.

Skew Bridge - A bridge in which the ends of the bridge are not square or perpendicular to the
centerline of the bridge.

Skew Crossing - Any crossing that is not perpendicular to the tracks.

Skin Friction - The friction between the outer surface of a pile or caisson and the surrounding
materials.

Slab - A flat, relatively thin, mass of wood, stone, concrete, or metal.

Slope Wall - A thin wall of concrete or of flat stones laid upon the face of a sloping bank of earth
to protect it from the erosive action of water.

Sole Plate - A plate riveted to the bottom flange of a plate girder to bear on the masonry plate.

Sounding - Measuring the depth of water.

Spacer Timber – See guard timber


Spalling - A surface deterioration of concrete resulting from several factors including moisture
damage, poor concrete, and reactive aggregates.

Span - The distance between two supports holding up a structure. The structure itself that rests on
the supports, as a span of a bridge.

Spandrel - The space from abutment to abutment in an arch bridge extending from the top of the
arch masonry to the top of the roadway.

Spider-rod - Steel rod that extends from the center casting of a rim bearing swing span through
each individual roller to hold a constant distance or diameter from the center. Also called radial
rods.

Splice Plate - A plate used in splicing or joining two parts of a member.

Spring Line - The line connecting the two opposite points where the curve of an arch begins.

Starling - A cutwater; the projecting end of a bridge-pier, usually so shaped as to allow ice, drift,
etc., to strike it without damage.

Stiffener - A secondary member, usually an angle, attached to a plate to prevent buckling.

Stone Masonry - A masonry structure constructed with stone.

Strain - The deformation per unit length caused by an external force applied to any piece of
material or to any bridge member.

Strand - One of the small threads or wires used in making rope.

Stress - An internal distributed force per unit area that resists the change in shape and size of a
body subjected to external forces.

Stringer - A longitudinal member extending from panel to panel of a bridge and supporting the ties
or the flooring.

Stringer Bracing - Diagonal bracing in the plane of the upper flanges of the stringers.

Strut - A bridge member carrying compression.

Sub-diagonal - A secondary member connecting the mid-point of a main diagonal with an adjacent
panel point.

Substructure - The part of any construction which supports the superstructure, such as piers and
abutments.
Superstructure - The part of a structure which receives the live load directly and carries the load
to the substructure.

Surcharge - The earth that lies both above and behind a retaining wall.

Surface - The vertical alignment of a section of track relative to its original survey.

Sway Bracing – (1) Bracing transverse to the planes of the trusses; used to resist wind pressure
and to prevent undue vibration. (2) Cross bracing in the plane of timber pile or frame bents.

Swing Bridge - A span that rotates about a vertical axis, so provide openings for the passage of
vessels.

T-beam - A reinforced concrete beam or a rolled structural section having a cross-section


resembling the letter "T."

Tangent - A straight line touching a curve at only one point. (may want to leave this out altogether)
The straight part of a railroad track.

Tension - The state or condition of being stretched.

Through Bridge - A bridge where a significant portion of the structure is above track level.

Through Girder - A type of structure where the support girders project above track level outside
of the tracks.

Through Plate Girder Span – A through span fabricated from steel plates and angles riveted,
bolted, or welded together where the sides of the girders come up above track level.

Through Truss - A truss that projects above track level and is braced across the top.

Tie Bolt - A round bolt with a square shank and lip for hooking ties to the flange of stringers.

Tie Spacer - A timber or steel strap that is connected to both ends of open deck bridge ties in order
to keep the ties evenly spaced. Also called a guard timber or spacer timber.

Toe - The foot of a slope. The front part of the base of an abutment or retaining wall.

Top Lateral Bracing - Lateral bracing in the plane of the top truss chords or beam/girder flanges.

Torque - The moment of a force or a system of forces tending to produce rotation.

Torsion - The twist or deformation of a body set up by a torque.

Tower - A vertical structure consisting of two or more bents of framework connected by bracing."

Tower Bracing - Bracing attached to the posts of a tower.


Tower Post - A member of a tower which carries load directly to the pedestal. A tower column.

Track Gage - The distance between the balls of the rails. (See Gage) Also the tool or device for
measuring or setting that distance.

Traction Bracing - Bracing in the plane of the bottom laterals which transfers the thrust of a
braking or accelerating train from the stringers to the trusses.

Transverse Beam - Any beam of a bridge that passes from one truss or girder to an adjacent truss
or girder.

Tread Plate - The bearing surface over which a wheel or roller moves.

Trestle - A bridge structure composed of bents or towers and supporting stringers or girders, which
may include a floor system.

Truss - A framed or jointed structure designed to act as a beam while each of its members is
primarily subjected to tension or compression stresses only.

Truss Pin - A pin used at the panel point of a truss to connect the several intersecting members.

Tunnel - An excavated passageway under the ground or the water.

Turntable - The framework under a swing span which transmits the load to the bearings. Also, a
stand alone structure used to rotate rolling stock or to line it up with approach tracks located around
the perimeter of the turntable pit.

Unsupported Length - The length of a compression member between the nearest points of lateral
restraint.

Viaduct - An extended bridge of many spans, mainly over dry ground. Usually consists of alternate
towers and open spaces or bays.

Wale - A flat piece of timber laid horizontally for bracing upright timbers and for guiding them
during driving, as in sheet piling.

Water Line - The intersection of the free surface of a body of water with any surface or object.

Web Plate - The plate forming the part of a girder between the top and bottom flanges.

Weep-hole - A hole in a wall for draining the water that tends to accumulate at the back.

Wind Loads - A load on a structure and train due to the pressure of the wind.
Wing Wall - One of the side walls of an abutment extending outward from the head wall in order
to hold back the slope of an embankment.

Wire Rope - A rope made of small strands of twisted wire often with a cotton or hemp center.

Worm - A helix or helical gear on a shaft which meshes into the worm gear.

Wrought Iron - In its perfect condition, wrought iron is simply pure iron, but, owing to impurities
(to a certain degree) being present, it only approximates to that condition.

Potrebbero piacerti anche