Download as pdf or txt
Download as pdf or txt
You are on page 1of 81

AS 5100.

2 Supp 1—2007
AP-G15.2C/06
AS 5100.2 Supp 1—2007

AS 5100.2 Supplement 1—2007


Accessed by Brandon & Associates refreshed 170613 on 23 Oct 2019 [OBSOLESCENT] (Document currency not guaranteed when printed)

Bridge design—Design loads—


Commentary
(Supplement to AS 5100.2—2004)
This Australian Standard Supplement was prepared by Committee BD-090, Bridge Design. It
was approved on behalf of the Council of Standards Australia on 26 October 2006.
This Supplement was published on 1 March 2007.

The following are represented on Committee BD-090:

• Association of Consulting Engineers Australia


• Australasian Railway Association
• AUSTROADS
• Bureau of Steel Manufacturers of Australia
• Cement Concrete & Aggregates Australia—Concrete
Accessed by Brandon & Associates refreshed 170613 on 23 Oct 2019 [OBSOLESCENT] (Document currency not guaranteed when printed)

• Engineers Australia
• Queensland University of Technology
• Steel Reinforcement Institute of Australia
• University of Western Sydney

Standards Australia wishes to acknowledge the participation of the expert individuals that
contributed to the development of this Standard through their representation on the
Committee and through public comment period.

Keeping Standards up-


up-to-
to- date
Australian Standards® are living documents that reflect progress in science, technology and
systems. To maintain their currency, all Standards are periodically reviewed, and new editions
are published. Between editions, amendments may be issued.

Standards may also be withdrawn. It is important that readers assure themselves they are
using a current Standard, which should include any amendments that may have been
published since the Standard was published.

Detailed information about Australian Standards, drafts, amendments and new projects can
be found by visiting www.standards.org.au

Standards Australia welcomes suggestions for improvements, and encourages readers to


notify us immediately of any apparent inaccuracies or ambiguities. Contact us via email at
mail@standards.org.au,
mail@standards.org.au or write to Standards Australia, GPO Box 476, Sydney, NSW 2001.
AS 5100.2 Supp 1—2007

AS 5100.2 Supplement 1—2007


Accessed by Brandon & Associates refreshed 170613 on 23 Oct 2019 [OBSOLESCENT] (Document currency not guaranteed when printed)

Bridge design—Design loads—


Commentary
(Supplement to AS 5100.2—2004)

First published as HB 77.2 Supp 1—1996.


Revised and redesignated AS 5100.2 Supp 1—2007.

COPYRIGHT
© Standards Australia
All rights are reserved. No part of this work may be reproduced or copied in any form or by
any means, electronic or mechanical, including photocopying, without the written
permission of the publisher.
Published by Standards Australia, GPO Box 476, Sydney, NSW 2001, Australia
ISBN 0 7337 8062 8
AS 5100.2 Supp 1—2007 2

PREFACE
This Commentary was prepared by the Standards Australia Committee BD-090, Bridge
Design to supersede HB 77.2 Supp 1, Australian Bridge Design Code—Design loads—
Commentary (Supplement to SAA HB 77.2—1996).
The objective of this Commentary is to provide users with background information and
guidance to AS 5100.2—2004.
The Standard and Commentary are intended for use by bridge design professionals with
demonstrated engineering competence in their field.
In this Commentary, AS 5100.2—2004 is referred to as ‘the Standard’.
The clause numbers and titles used in this Commentary are the same as those in AS 5100.2,
except that they are prefixed by the letter ‘C’. To avoid possible confusion between the
Commentary and the Standard, a Commentary clause is referred to as ‘Clause C…..’ in
Accessed by Brandon & Associates refreshed 170613 on 23 Oct 2019 [OBSOLESCENT] (Document currency not guaranteed when printed)

accordance with Standards Australia policy.


3 AS 5100.2 Supp 1—2007

CONTENTS

Page
C1 SCOPE AND GENERAL ........................................................................................... 4
C2 REFERENCED DOCUMENTS.................................................................................. 4
C3 DEFINITIONS............................................................................................................ 4
C4 NOTATION................................................................................................................ 4
C5 DEAD LOADS ........................................................................................................... 5
C6 ROAD TRAFFIC ........................................................................................................ 7
C7 PEDESTRIAN AND BICYCLE-PATH LOAD ........................................................ 26
C8 RAILWAY TRAFFIC............................................................................................... 26
C9 MINIMUM LATERAL RESTRAINT CAPACITY .................................................. 34
Accessed by Brandon & Associates refreshed 170613 on 23 Oct 2019 [OBSOLESCENT] (Document currency not guaranteed when printed)

C10 COLLISION LOADS ............................................................................................... 34


C11 KERB AND BARRIER DESIGN LOADS AND OTHER REQUIREMENTS
FOR ROAD TRAFFIC BARRIERS ......................................................................... 37
C12 DYNAMIC BEHAVIOUR........................................................................................ 40
C13 EARTH PRESSURE................................................................................................. 42
C14 EARTHQUAKE FORCES........................................................................................ 44
C15 FORCES RESULTING FROM WATER FLOW ...................................................... 49
C16 WIND LOADS ......................................................................................................... 54
C17 THERMAL EFFECTS .............................................................................................. 57
C18 SHRINKAGE, CREEP AND PRESTRESS EFFECTS ............................................. 61
C19 DIFFERENTIAL MOVEMENT OF SUPPORTS ..................................................... 62
C20 FORCES FROM BEARINGS................................................................................... 63
C21 CONSTRUCTION FORCES AND EFFECTS.......................................................... 64
C22 LOAD COMBINATIONS ........................................................................................ 64
C23 ROAD SIGNS AND LIGHTING STRUCTURES .................................................... 65
C24 NOISE BARRIERS .................................................................................................. 67

APPENDIX CA DESIGN LOADS FOR MEDIUM AND SPECIAL


PERFORMANCE LEVEL BARRIERS .................................................... 68
AS 5100.2 Supp 1—2007 4

STANDARDS AUSTRALIA

Australian Standard
Bridge design—Design loads—Commentary
(Supplement to AS 5100.2—2004)

C1 SCOPE AND GENERAL


C1.1 Scope
(No Commentary)
C1.2 General
Although details of loads commonly occurring on bridge structures are outlined in the
Accessed by Brandon & Associates refreshed 170613 on 23 Oct 2019 [OBSOLESCENT] (Document currency not guaranteed when printed)

Standard, the designer should consider the possibility of other unusual loads occurring. The
general principles of AS 5100.1, Bridge design—Scope and general principles, should be
observed when assessing unusual loads, and most importantly the designer should ensure
that damage cannot occur which is out of all proportion to the original cause.
It is particularly important that the fundamental design information, including as-
constructed data, be recorded on the front sheet of the bridge drawings.
The abbreviation of SM1600 is introduced to provide a single abbreviation to indicate that
the bridge has been designed for the worst effects induced by each of the W80, A160,
M1600 and S1600 road traffic design loads.
The abbreviation 300LA is introduced to provide a single abbreviation to indicate that the
bridge has been designed for the worst effects induced by each of the 360 kN axle load, the
1560 kN simulated locomotive and the 1560 kN simulated locomotive coupled to any
number of 1200 kN vehicles as specified in Clause 8 of the Standard.

C2 REFERENCED DOCUMENTS
The Standards listed in Clause 2 are subject to revision from time to time and the current
edition should always be used. The currency of any Standard may be checked with
Standards Australia.

C3 DEFINITIONS
Technical definitions are provided in the Clause. Some technical definitions that are
applicable to only one Clause are given in the Clause in which they are relevant.

C4 NOTATION
The basis of the notation is generally in accordance with ISO 3898, Bases for Design of
Structures—Notations—General Symbols. Standards Australia’s policy is to use ISO
recommendations on notation wherever practicable in structural design Standards such as
AS/NZS 1170 series, AS 2327.1, Composite structures—Simply supported beams, AS 3600,
Concrete structures, AS 4100, Steel structures and AS/NZS 4600, Cold-formed steel
structures.

 Standards Australia www.standards.org.au


5 AS 5100.2 Supp 1—2007

C5 DEAD LOADS
C5.1 General
The weights per cubic metre given in Tables C5.1(A) and C5.1(B) may be used in
calculating the nominal dead load unless a more precise determination has been made. The
value to be used should be the mean value, not a characteristic value, as the load factors
take this into account. Where a range of values is given, calculations should be performed
using the extremes of the expected range and the most critical case used for design.
The value of g, which has been used in calculating the weights, per cubic metre, given in
the tables is 9.8 m/s 2.
Approximate values of concrete weight per cubic metre are given in Table C5.1(B) for
different aggregate types and cement contents.

TABLE C5.1(A)
WEIGHTS FOR CALCULATING
Accessed by Brandon & Associates refreshed 170613 on 23 Oct 2019 [OBSOLESCENT] (Document currency not guaranteed when printed)

NOMINAL DEAD LOADS

Weight per cubic metre


Material
kN/m 3

Aluminium alloy 26.7


Bituminous wearing surface, asphalt 22.0
Compacted earth filling 16.0–19.0
Compacted gravel, road metal 19.0–23.0
Concrete (light weight) 22.5–26.0
Masonry 23.5
Neoprene 11.3
Sand-fine (dry) 15.5–17.5
Sand-coarse (dry) 18.0–19.5
Sand (saturated) 22.5
Steel and other ferrous metals 77.0
Timber, softwood (see Note) 8.3
Timber, hardwood (see Note) 12.3
Water, fresh 9.8
Water, salt 10.0
NOTE: AS 1684.1, Residential timber-framed construction—
Design criteria.

www.standards.org.au  Standards Australia


AS 5100.2 Supp 1—2007 6

TABLE C5.1(B)
WEIGHT PER CUBIC METRE OF UNREINFORCED CONCRETE
Density of coarse Cement Weight per
Typical coarse aggregates aggregates content cubic metre
kg/m 3 kg/m 3 kN/m 3
Adelaide Brisbane Perth Sydney 450 24.0
quartzite gravel granite gravel
{ 2500 330
450
22.5
24.5

{
2700
Melbourne Sydney 330 23.0
basalt basalt 450 25.5

{
2900
330 24.0
Hobart 450 26.0
3100
dolerite 330 25.0
NOTES:
Accessed by Brandon & Associates refreshed 170613 on 23 Oct 2019 [OBSOLESCENT] (Document currency not guaranteed when printed)

1 The values given in the Table apply to normal concrete, have no added air and the accuracy is
approximately ±0.5 kN/m 3 .
2 The values given do not include any allowance for reinforcement. For reinforced concrete, the values
should be increased by 0.6 kN/m 3 for each one percent by volume of reinforcement.

C5.2 Dead load of structure


The load factors to be applied to concrete dead loads are based on the assumption that the
designer has sufficient knowledge of the concrete properties to select a design density close
to that of the concrete used in the structure. Where the designer has no knowledge of the
concrete density, a suitably high value should be selected for calculations where the
concrete dead load is unfavourable, and a low value for calculations where the dead load
increases the structural safety.
A balanced cantilever is a statically determinate structure in which the cantilevers are
rigidly connected to and extend either side of the support, such that the dead load bending
moments at either side of the support are approximately balanced.
An anchor cantilever is a statically determinate structure, which is supported at two points
and extends past at least one of these supports to form a cantilever.
For either of these structural types, or other similar types, and for certain elements of the
structure which are subject to both unfavourable and favourable dead loads, such as the
supports, the design dead loads should be obtained by applying the load factors, given in
Table 5.2, for type structures (b) or (c), to the dead load of the appropriate parts of the
structure.
For other structural elements, which are for instance subject to unfavourable dead loads
only such as the cantilevers themselves, the design dead loads should be obtained by
applying the load factors given in Table 5.2, type structure (a), to the nominal dead load on
the element.
C5.3 Superimposed dead load
For the design of a bridge having a concrete wearing surface, the design load due to an
added 75 mm thick bituminous wearing surface should be considered even though there
may be no immediate intention to place such a layer.
The design superimposed dead load should be obtained by applying the load factors given
in Table 5.3 of the Standard to the nominal superimposed dead load on the element.

 Standards Australia www.standards.org.au


7 AS 5100.2 Supp 1—2007

C5.4 Soil loads on retaining walls and buried structures


The Commentary to AS 5100.3, Bridge design—Foundation and soil supporting structures
contains design information on various soils and rocks, their properties and methods of
calculating soil imposed loads and resistances.
Soil loads and properties may be also obtained from AS 4678, which should be treated as a
guide only. The design methods and factors specified in AS 4678 are not compatible with
the design methods and factors contained within AS 5100.3. Nevertheless, AS 4678 does
contain much information which is useful for design purposes.
The factors in Table 5.4 of the Standard account for the differing levels of uncertainty
associated with controlled fills and other types of fills and soils. These factors apply to
active, at rest and passive conditions, as relevant.
Pressures imposed by groundwater can be significant, with effects depending on its level
with respect to the structure and the effectiveness of the drainage provisions. Variations in
groundwater levels should be accounted for in the design.
Accessed by Brandon & Associates refreshed 170613 on 23 Oct 2019 [OBSOLESCENT] (Document currency not guaranteed when printed)

C5.5 Railway ballast and track loads


The maximum amount of ballast to be considered should be the amount that might be
placed on a flat deck with normal height kerbs where excess ballast would fall to the ground
below the bridge. It does not apply to through girders, where the maximum amount of
ballast should be taken as the maximum amount that could reasonably be foreseen.
The design superimposed dead load for railway bridges should be obtained by applying the
load factors given in Table 5.5 of the Standard to the nominal railway ballast and track
loads on the element.
A typical ballasted track structure consistent with the 300LA live load will have a dead load
of approximately 12 kPa. The dead load for a transom top structure will be approximately
5 kN/m. It should be noted that axle loads are not the only factors to be considered in the
design of track work, e.g., the track structure for a high speed passenger railway may be
identical to that for a heavy haul railway because of track stability considerations.

C6 ROAD TRAFFIC
C6.1 General
The Austroads Bridge Design Code, ABDC-1992 (HB 77.2—1996) (Ref. 1) loading model
was revised in 1999 with the objective of providing for the maximum anticipated increases
in vehicle size and weight over the life of bridges. Thus the loading model is based on the
maximum anticipated traffic load well into the future rather than the traffic load that exists
at the time of construction.
Economic studies (see Gordon and Boully (1997) (Ref. 2)) have shown that marginal costs
associated with providing stronger bridges are small in comparison with the benefits of a
more efficient transport system. It is intended that bridges designed for this load will have
adequate strength for their expected lifespan of at least 100 years.
The traffic loads specified in the Standard, including the various factors, have been
determined to cover the effects of—
(a) a large number of legal or marginally overloaded vehicles. This defines the loading
spectra;
(b) a small number of grossly overloaded vehicles; and
(c) heavily loaded vehicles following one another.

www.standards.org.au  Standards Australia


AS 5100.2 Supp 1—2007 8

The loads applied by heavy vehicles such as cranes, off-road machinery and abnormal
heavy indivisible loads carried by multi-axle platforms that travel by roads on permits are
special cases. The effects induced by these vehicles, taking into consideration appropriate
load factors, dynamic load allowance and widths, will be generally smaller than the design
loads.
The load model developed should ensure bridges constructed in accordance with the
Standard will not become impediments to future productivity enhancements in road
transport. It was adopted after the consideration of —
(i) the freight task as a function of the density of the freight;
(ii) the costs and benefits associated with heavier and longer vehicles;
(iii) the likely limits associated with vehicle technology and safety issues;
(iv) the environmental consequences of larger vehicles; and
(v) the cost of upgrading existing bridges to the Standard.
Accessed by Brandon & Associates refreshed 170613 on 23 Oct 2019 [OBSOLESCENT] (Document currency not guaranteed when printed)

The load model [see Gordon and Boully (1998) (Ref. 3) and Heywood, Gordon and Boully
(2000) (Ref. 4)] is designed to induce effects in bridges equivalent to the average extreme
daily effects, i.e., the average of the largest events each day, induced by the traffic stream.
The load models are presented for a single lane of traffic. The single lane models are
combined using accompanying lane factors to represent multi-lane traffic. The load models
correspond to the following three segments of loading (see Figure C6.1):
(A) Slightly overloaded individual axles and axle groups moving over bridges. AB as
shown in Figure C6.1 represented by the W80 wheel load, the A160 axle load and the
individual axle groups within the M1600 moving traffic load.
(B) Legally loaded groups of axles or entire vehicles moving over bridges where the
distance between axle groups is at a minimum. BCD shown in Figure C6.1
represented by the M1600 moving traffic load.
(C) Queues of stationary vehicles. DE shown in Figure C6.1 represented by the S1600
stationary traffic load. Note that point E is approximately 60% of point F that
corresponds to the load per unit length of a 200 m long queue of the maximum
anticipated future legally loaded vehicles.
The M1600 and S1600 load models are relatively long. They are a response to the frequent
observations of heavy vehicles travelling with little distance between each other and the
widespread application of combination vehicles such as truck-trailers, B-doubles and road
trains throughout Australia. B-triple trials and advances in intelligent transport systems
reinforce the need for design vehicles that represent multiple or combination vehicles
irrespective of the vehicle type permitted to use the route.
Note that the first letter of the load model infers the traffic condition being modelled and
the numbers indicate the total weight (approximately) in kilonewtons of the wheel/axle or
the truck plus lane load at a loaded length of 25 m.

 Standards Australia www.standards.org.au


9 AS 5100.2 Supp 1—2007

FIGURE C6.1 SUMMARY OF LOAD MODEL

C6.2 SM1600 loads


Accessed by Brandon & Associates refreshed 170613 on 23 Oct 2019 [OBSOLESCENT] (Document currency not guaranteed when printed)

C6.2.1 W80 wheel load


The W80 wheel and A160 axle loads provide a minor increase in the factored serviceability
limit state loading and approximately the same factored ultimate limit state loading as the
corresponding loads specified in ABDC-1992 (HB 77.2—1996) (Ref. 1).
The W80 wheel and A160 axle loads ensure local elements are sufficiently strong to
support an overloaded wheel or axle. It is somewhat heavier than the anticipated future axle
loads as studies have shown that individual axles are the most likely to be overloaded.
The W80 wheel load ensures that substantial wheel loads are applied at any point of the
road surface. This ensures that bridge decks are adequate for standard heavy vehicles,
special purpose vehicles such as low loaders, cranes and agriculture machinery.
The tyre footprint and track are consistent with new vehicle technologies that are trending
towards wider tracks and higher tyre pressures.
C6.2.2 A160 axle load
(For Commentary, see Clause C6.2.1)
C6.2.3 M1600 moving traffic load
The M1600 moving and S1600 stationary traffic load models are designed to induce effects
equivalent to the moving traffic stream and queues of stationary traffic. In particular, this
envisages rigid, truck trailer, articulated, B-double, B-triple and road train heavy vehicles.
The following provides an indication of representative future vehicles that served as the
basis for the SM1600 design traffic load. In the development of SM1600, two families of
vehicles were considered:
(a) Vehicles of similar length to existing freight vehicles operating at a freight density of
about 0.58 t/m 3 .
(b) Vehicles with shortened wheelbases for operation on routes with restricted swept
paths or for transporting heavy materials operating at a freight density of about
0.73 t/m 3 .
Tables C6.2(A) and C6.2(B) provide details of representative vehicles in each of the above
categories. It is envisaged that future vehicles will have increased numbers of axles to
minimize effects on pavements and, hence, actual vehicles will probably use tri-axle drive
groups and quad-axle trailer groups.
For comparison purposes, Table C6.2(C) provides details of corresponding current vehicle
configurations and current freight density of about 0.28 t/m3 .
www.standards.org.au  Standards Australia
AS 5100.2 Supp 1—2007 10

TABLE C6.2(A)
CURRENT LENGTH VEHICLES—FUTURE FREIGHT DENSITY = 0.58 t/m 3
O/A Gross mass
Configuration, axle distances and axle group mass
Vehicle length
(see Note)
(m) (tonnes)
o oo ooo
Articulated 19.0 74 1.6 m 6.4 m 15.9 m
7t 27 t 40 t
o oo ooo ooo
B-double 25.0 114 1.6 m 5.6 m 13.6 m 22.6 m
7t 27 t 40 t 40 t
o oo ooo ooo ooo
B-triple 33.0 154 1.6 m 5.6 m 13.6 m 21.6 m 30.6 m
Accessed by Brandon & Associates refreshed 170613 on 23 Oct 2019 [OBSOLESCENT] (Document currency not guaranteed when printed)

7t 27 t 40 t 40 t 40 t
o oo ooo oo ooo oo ooo
RT triple 53.5 206 1.6 m 6.4 m 15.9 m 23.8 m 33.3 m 41.2 m 50.7 m
7t 27 t 40 t 27 t 40 t 27 t 40 t
NOTE: Distances are from front of vehicle to centre-line of axle group.

TABLE C6.2(B)
SHORT VEHICLES—FUTURE FREIGHT DENSITY = 0.73 t/m 3
O/A Gross mass
Configuration, axle distances and axle group mass
Vehicle length
(see Note)
(m) (tonnes)
o oo ooo
Articulated 14.8 73 1.25 m 5.25 m 12.8 m
7t 26.5 t 39.5 t
o oo ooo ooo
B-double 21.4 112.5 1.25 m 5.25 m 11.7 m 19.2 m
7t 26.5 t 39.5 t 39.5 t
o oo ooo ooo ooo
B-triple 27.9 152 1.25 m 5.25 m 11.7 m 18.2 m 25.7 m
7t 26.5 t 39.5 t 39.5 t 39.5 t
o oo ooo oo ooo oo ooo
RT triple 40.6 205 1.25 m 5.25 m 12.8 m 18.1 m 25.6 m 31 m 38.5 m
7t 26.5 t 39.5 t 26.5 t 39.5 t 26.5 t 39.5 t
NOTE: Distances are from front of vehicle to centre-line of axle group.

 Standards Australia www.standards.org.au


11 AS 5100.2 Supp 1—2007

TABLE C6.2(C)
CURRENT LENGTH VEHICLES—CURRENT FREIGHT DENSITY = 0.28 t/m3
O/A
Gross mass Configuration, axle distances and axle group mass
Vehicle length
(tonnes) (see Note)
(m)
o oo ooo
Articulated 19.0 42.5 1.6 m 6.4 m 15.9 m
6t 16.5 t 20 t
o oo ooo ooo
B-double 25.0 62.5 1.6 m 5.6 m 13.6 m 22.6 m
6t 16.5 t 20 t 20 t
o oo ooo ooo ooo
B-triple 33.0 82.5 1.6 m 5.6 m 13.6 m 21.6 m 30.6 m
Accessed by Brandon & Associates refreshed 170613 on 23 Oct 2019 [OBSOLESCENT] (Document currency not guaranteed when printed)

6t 16.5 t 20 t 20 t 20 t
o oo ooo oo ooo oo ooo
RT triple 53.5 115.5 1.6 m 6.4 m 15.9 m 23.8 m 33.3 m 41.2 m 50.7 m
6t 16.5 t 20 t 16.5 t 20 t 16.5 t 20 t
NOTE: Distances are from front of vehicle to centre-line of axle group.

A series of four tri-axle groups and a uniformly distributed load have been chosen as the
base elements of both the M1600 and S1600 traffic load models. This style of model was
selected because it is simple, resembles heavy vehicles and research has shown that a series
of axles is the best way to model both the bending and the shear effects induced by actual
traffic.
The M1600 moving traffic load models moving traffic whereas the S1600 stationary traffic
load models queues of stationary traffic. This division allows for the logical application of
dynamic, braking and centrifugal effects to the M1600 moving traffic load but not to the
S1600 stationary traffic load.
Each traffic load model is equivalent to either one heavy combination vehicle or two heavy
vehicles in the same lane, together with an accompanying stream of general traffic. Four tri-
axle groups plus a uniformly distributed load equal to the length of the vehicle represents
the heavy vehicle. The extension of the uniformly distributed load at either end of the heavy
vehicle represents the accompanying traffic stream. Note that the total load associated with
the heavy vehicle component of both the load models is approximately the same (1600 kN).
The effects induced in bridge elements are a function of the—
(i) mass of the vehicle;
(ii) the vehicle configuration;
(iii) the distribution of mass between axle groups; and
(iv) the bridge configuration, simply supported or continuous, and material properties.
Generally, critical events result from vehicles that are both short and heavy. The truck
component of the M1600 and S1600 traffic loads correspond to the maximum anticipated
future legally loaded B-triples or equal, operating with minimum distances between axles.
Note that at these axle loads, approximately 80% of the freight task is expected to be
limited by volume restrictions rather than weight. Consequently, most heavy vehicles will
operate with distances between axle groups that are greater than the minimum.

www.standards.org.au  Standards Australia


AS 5100.2 Supp 1—2007 12

The traffic load model features a range of distances between axle groups as the spacing
between axle groups required to induce the critical effects varies with span and
configuration (simply supported versus continuous). The minimum spacing between tri-axle
groups (3.75 m) provides a balance between the needs of the transport industry and the
provision of bridges. The variable spacing between the second and third tri-axle groups
models following vehicles. This is especially significant for continuous bridges.
The S1600 stationary traffic load was derived from simulation studies based on
weigh-in-motion records for heavily trafficked routes. The simulation studies showed that
the average load per unit length reduced with length in a manner that was consistent with a
heavy combination vehicle accompanied by a uniformly distributed load.
The magnitude of the S1600 uniformly distributed load takes into account the space
between vehicles, the mix of cars and partially loaded heavy vehicles. As a consequence,
the total load for a 200 m loaded length is approximately 60% of a queue of the maximum
anticipated future legally loaded B-triples.
The trailing uniformly distributed load associated with the M1600 moving lane load is
Accessed by Brandon & Associates refreshed 170613 on 23 Oct 2019 [OBSOLESCENT] (Document currency not guaranteed when printed)

exactly 1/4 of the trailing uniformly distributed load associated with the S1600 stationary
traffic load. The difference is a consequence of the lower densities associated with moving
traffic compared to stationary traffic. It is based on the assumption that the spacing between
vehicles increases by 1 m for every 2 km/h (see Ontario Highway Bridge Design Code,
1983) (Ref. 5). As a consequence, the moving traffic load is a function of speed (see
Figure C6.2.3); however, the single heavy vehicle dominates the load with speed being a
secondary effect. In order to simplify the load provisions, a single load for all speeds has
been adopted.
The M1600 moving lane load is relevant to all spans in that it forms the basis for
determining the co-existing vertical loads and braking, or centrifugal forces. The S1600
traffic load becomes critical once the effects of its heavier loads are greater than the
dynamic effects plus the M1600 traffic load.

FIGURE C6.2.3 INFLUENCE OF SPEED ON LANE LOAD

C6.2.4 S1600 stationary traffic load


(For Commentary, see Clause C6.2.3)

 Standards Australia www.standards.org.au


13 AS 5100.2 Supp 1—2007

C6.3 Heavy load platform


The W80, A160, M1600 and S1600 traffic load models, generally cover the effects induced
by the HLP320 and HLP400 specified in ABDC-1992 (HB 77.2—1996) (Ref. 1) because
these HLP design loads are assumed to occupy two design lanes. Consequently, no further
provision for heavy load platform loading need be made unless specified by the authority.
Heavy load platform loading may control the design of some structures such as single lane
freeway ramps on a separate structure and buried structures. It may also control
serviceability limit state design for some structures, including the crack control stress
increments specified in AS 5100.5.
Should the authority specify a heavy load platform, the lateral position, tolerance on
position, load factors and restrictions on accompanying traffic also need to be specified.
The position of the heavy load platform loading and any other traffic loading assumed in
the design should be clearly shown on the drawings.
C6.4 Tramway and railway loads
Accessed by Brandon & Associates refreshed 170613 on 23 Oct 2019 [OBSOLESCENT] (Document currency not guaranteed when printed)

Tramway and railway design loads and load factors should be obtained from the appropriate
authority. These loads do not have to be considered in conjunction with heavy load platform
loads.
C6.5 Number of lanes for design and lateral positioning
The Clause is not meant to penalize bridges with wide shoulders for traffic safety. The
authority should assess whether it is feasible or practical to use the bridge with the
calculated number of standard design lanes if this number is greater than the number of
lanes marked on the bridge. Bridges in urban areas should generally be designed for the
maximum number of standard design lanes possible since future local traffic conditions are
difficult to predict and marked traffic lanes could be altered. Determination of the number
of design lanes should also consider the possibility that in the future a walkway may be
converted to a roadway. This possibility is particularly significant with the SM1600 traffic
load model as it is significantly heavier than the pedestrian load.
C6.6 Accompanying lane factors
The accompanying lane factors given in Table 6.6 take into account the reduced probability
that extreme loads will occur simultaneously in all lanes. These factors are further modified
to account for the fact that dynamic load effects will also be reduced because all vehicles on
the bridge will not produce vibration effects exactly in phase. The errors induced by
applying this further reduction to the accompanying load factor rather than modifying the
dynamic load allowance, are not significant.
The accompanying lane model is based on the concept that extreme events are accompanied
by more common events. As the number of lanes increases, the number and size of the
accompanying vehicles reduces. This phenomenon is observed in simulation studies even
when the distribution of heavy vehicles is the same in all lanes. The accompanying lane
model introduces greater eccentricities into the load compared with the averaging effect
associated with the lane reduction factor approach. Single box girders, cantilevered
headstocks, through trusses and the like are examples of structures that experience
increased effects resulting from eccentric loads.
The extreme heavy vehicle can occur in any lane. The standard design lane numbers should
be varied to induce the greatest effect in each element considered. Some examples are
shown in Figure C6.6. The critical arrangement will vary from element to element. There
are a large number of possible combinations that can usually be reduced with experience.

www.standards.org.au  Standards Australia


AS 5100.2 Supp 1—2007 14

FIGURE C6.6 ACCOMPANYING LANE MODEL—EXAMPLES SUPERIMPOSED


ON A BRIDGE CROSS-SECTION

C6.7 Dynamic load allowance


C6.7.1 General
The dynamic interaction between road profiles, vehicle speed and characteristics, and
bridges is complex. Also the dynamic and static responses of bridges are not the same. The
dynamic load allowance (α) is an equivalent static load, expressed as a fraction of the
design traffic loading, which is considered for design purposes to be equivalent to the
Accessed by Brandon & Associates refreshed 170613 on 23 Oct 2019 [OBSOLESCENT] (Document currency not guaranteed when printed)

dynamic and vibratory effects of the interaction of the moving load and the bridge. See
Clough and Penziem, 1975 (Ref. 6).
The Clause has been modified to reflect the changes in the Canadian Highway Bridge
Design Code 2000 (Ref. 7). It represents a simplification in the application of the
theoretical work on which the provisions of the Ontario Highway Bridge Design Code
(OHBDC 1983) (Ref. 5) and the ABDC-1992 (HB 77.2—1996) (Ref. 1) were based.
It also takes into consideration the findings of more than 80 field investigations into the
dynamic response of bridges in Australia and New Zealand and related theoretical analysis,
see Heywood and Boully (2003) (Ref. 8) and Heywood (2000) (Refs 9 and 10).
C6.7.2 Magnitude
In an analogous manner to the Canadian Highway Bridge Design Code 2000 (Ref. 7),
different dynamic load allowance values are specified for individual axles, single axle
groups and more than one axle group of the traffic design load vehicles. The dynamic load
allowance specified for the heavy load platform loading remains unchanged. It is applicable
to HLP loadings travelling at up to 10 km/h.
Research has shown that higher values of dynamic load allowance are required for short
span bridges where frequency matching between the structure and individual axles or single
axle groups may induce axle hop. Provisions in the Ontario Highway Bridge Design Code
1983 (Ref. 5) and the ABDC-1992 (HB 77.2—1996) (Ref. 1) were based on tests that did
not adequately cover short span bridges. See Green (1977) (Ref. 11).
The Standard, like the Canadian Highway Bridge Design Code 2000 (Ref. 7), has simplified
the design approach by specifying dynamic load allowance as a function of the number of
axles or axle groups creating the critical action in the component under consideration.
The A160 axle and the single M1600 tri-axle are generally critical for short span lengths
where the natural frequency is generally greater than 10 Hz. This is the frequency range that
was omitted from the previous provisions and shown to be important by local and overseas
research.
The maximum dynamic load allowance of 0.4 specified for the single W80 wheel load and
A160 axle load reflects responses observed in short span structures and individual
components.

 Standards Australia www.standards.org.au


15 AS 5100.2 Supp 1—2007

The dynamic load allowance of 0.35 specified for a single tri-axle of M1600 moving traffic
load again reflects the response observed in short span structures. The marginally reduced
value compared to that for a single axle allows for some interaction and the spread of load
between axles within the group. It is marginally higher than the value of 0.3 applied by the
Canadian Code to any two axles and the prime-mover of their 5 axle articulated design
vehicle. The Standard allows for rougher roads than the Canadian Code as demonstrated by
local research. See Prem and Heywood (2000) (Ref. 12).
The single dynamic load allowance of 0.3 to be applied to the actions induced by two or
more axle groups of the M1600 moving traffic load is less than the maximum value of 0.4
specified in the ABDC-1992 (HB 77.2—1996) (Ref. 1) for natural frequencies of about 2.5
to 4.5 Hz and span lengths of about 22 to 40 m. It is representative of typical, but not
extreme, values obtained by local testing for natural frequencies less than 10 Hz and span
lengths greater than 10 m. Again this value is marginally higher than the value of 0.25
specified to be applied to more than two axles by the Canadian Highway Bridge Design
Code 2000 (Ref. 7).
Accessed by Brandon & Associates refreshed 170613 on 23 Oct 2019 [OBSOLESCENT] (Document currency not guaranteed when printed)

Local and overseas research has shown that the magnitude of the dynamic load allowance
generally reduces as the mass of the vehicle increases. Hence, it is not appropriate to factor
ultimate limit state actions, induced by heavy design vehicles, by the maximum, measured
dynamic load allowances from partly loaded trucks, which commonly exceed the design
values given in the Standard. It is more appropriate to factor serviceability limit state
actions by the maximum values but no differentiation is made in the Standard because of
the large number of other variables and complex interactions involved. The reduced
dynamic load allowance for two or more axle groups reflects the interaction of the axle
groups of one or more vehicles traversing the bridge.
It is now recognized that for multi-lane bridges, there is a reduced probability that the
maximum dynamic load effect will occur simultaneously in all loaded lanes. The
accompanying load factors given in Table 6.6 of the Standard take this into consideration
(see also Clause C6.6).
For the special case of the design of expansion joints, the dynamic load allowance is
specified in AS 5100.4; the dynamic load allowance to be used for the design of manhole
covers, grates and the like are specified in AS 3996.
C6.7.3 Application
The dynamic load allowance (α) is only applied to the moving loads, i.e., W80, A160 and
M1600. The dynamic load allowance is not applied to the S1600 stationary traffic load
because it does not induce any dynamic response in the bridge.
No dynamic load allowance is applied to M1600 moving traffic load when determining
braking forces (F BS) and centrifugal forces (F c), as these forces are generated by
accelerating the vehicle mass parallel to the plane of the deck.
The design of arch type buried structures is generally controlled by the loading from a
single axle or single axle group. The dynamic load allowance values to be used for these
loadings make some allowance for any local irregularities in the road profile in the vicinity
of the crown of the arch due to settlement of the backfill material. Field investigations
indicate that the dynamic effects of loads decrease with increasing depth. A dynamic load
allowance of 0.1 to be applied to buried structures with more than 2 m of cover reflect the
high energy absorption capacity of the soil mass.
C6.7.4 Dynamic load reversal
Bridges continue to vibrate after the passing of vehicles and this can reduce the minimum
reaction or even cause uplift. By neglecting the attenuation of the dynamic response it
follows that the magnitude of this effect is approximately the same as the dynamic load
allowance.

www.standards.org.au  Standards Australia


AS 5100.2 Supp 1—2007 16

This phenomenon is suspected as being a contributory cause of apparent low frictional


resistance of bearings, the walking of bearings, fatigue failures of metal components of
deck joints and also of flexural cracks on top of light prestressed concrete beams. The
designer should give special attention to detailing where such events are likely to occur.
C6.8 Horizontal forces
C6.8.1 Centrifugal forces
Transverse forces are required in order for vehicles to move around a curve. These forces
are referred to as centrifugal forces (Fc). The magnitude of Fc depends on the mass (m) and
speed (v) of the vehicle and the radius (r) of the curve (F c = mv2/r). Geometric
considerations and the road operating speed determine two of the parameters leaving the
mass to be specified. The M1600 moving lane load defines the mass and its distribution
along the length of a structure. Thus the co-existing gravitational and centrifugal effects are
defined for those occasions where the combination of vertical and horizontal forces need to
be considered.
Accessed by Brandon & Associates refreshed 170613 on 23 Oct 2019 [OBSOLESCENT] (Document currency not guaranteed when printed)

The centrifugal force acts through the centre of gravity of the vehicle. To simplify the
design process, the Standard allows the centrifugal force to be applied at the road surface.
The centrifugal acceleration is limited to (0.35 + θ)g for vehicle stability reasons. At the
ultimate limit state, Fc has a load factor of 1.8 applied. Thus the ultimate limit state is
equivalent to a grossly overloaded vehicle experiencing a lateral acceleration of 0.35g or a
combination vehicle cornering with an acceleration of 1.8 × 0.35g = 0.63g (assumes θ = 0).
This is consistent with the upper limit of lateral acceleration for roll-over of low centre of
gravity legally loaded vehicles which is approximately 0.5g.
C6.8.2 Braking forces
An accelerating or decelerating traffic stream induces longitudinal forces that are referred
to as braking forces. In general, low rates of braking are a regular in-service condition but
provision also needs to be made for extreme events.
The maximum braking force achieved is a function of the—
(a) limiting coefficient of friction between the tyres and the pavement times the weight of
the vehicle;
(b) capacity of the braking and suspension systems fitted to the vehicle;
(c) behaviour of the driver; and
(d) number of vehicles braking.
At a certain load level, the capacity of the braking system and the friction between the tyres
and the pavement are approximately equal. This corresponds to the maximum braking force
that can be achieved, even for higher load levels. Consequently, a grossly overloaded
vehicle is unable to apply a braking force which is higher than the capacity of the braking
system, i.e., a highly overloaded truck will decelerate at a slower rate than a vehicle at its
legal load.
The serviceability limit state loading corresponds to a relatively infrequent event. This
ensures that braking does not cause damage that requires regular maintenance to bearings,
joints and substructure elements. The ultimate limit state event would be expected to cause
damage that may require repair.

 Standards Australia www.standards.org.au


17 AS 5100.2 Supp 1—2007

The Standard considers the following limiting scenarios:


(i) Single vehicle stopping Braking tests conducted in Australia have shown the
following performance:
(A) Six-axle artics .......................... maximum decelerations of 0.75g at legal loads.
(B) B-triples ..................................... maximum decelerations of 0.6g at legal loads.
(C) Overloaded trucks ......... maximum decelerations reduce with increasing weight.
The Australian design rules require braking systems to be capable of decelerating
heavy vehicles at a minimum rate of approximately 0.45g. The braking force (F BS) is
based on this deceleration rate. The upper limit of 720 kN corresponds to 0.45 times
the weight of the vehicle component of the M1600 moving traffic load.
The ultimate limit state braking forces that have been adopted correspond to a range
of mass and deceleration rates. Two examples are as follows:
(1) One legally loaded combination vehicle (or equal) braking at
Accessed by Brandon & Associates refreshed 170613 on 23 Oct 2019 [OBSOLESCENT] (Document currency not guaranteed when printed)

1.8 × 0.45g = 0.81g.


(2) One grossly overloaded combination vehicle (or equal) braking at 0.45g.
The single vehicle stopping scenario will be critical for most bridges.
(ii) Multi-lane moving traffic stream stopping The forces applied by a braking traffic
stream are critical for bridges with large road surface areas between expansion joints.
A deceleration of 0.15g is applied to the M1600 moving traffic load to generate the
serviceability limit state braking force. This is consistent with deceleration rates
adopted in road design. The ultimate limit state for the moving traffic stream
corresponds to a braking deceleration of 0.27g.
Figure C6.8.2 compares the ultimate limit state braking force due to a single vehicle
stopping (F BS) with those for the multi-lane traffic stream stopping (F BM). For typical
lengths between expansion joints, the emergency braking of a single M1600 vehicle
dominates.
The braking load has been expressed as a portion of the M1600 moving traffic load. This is
consistent with the basic physics. It provides a framework for the designer to distribute the
braking forces longitudinally and transversely in accordance with the distribution of load
within the M1600 moving traffic load. It also models the co-existing vertical and horizontal
effects induced by the traffic stream stopping should a structure be sensitive to these
combined effects. The braking force is to be applied in some or all lanes as necessary. This
allows the twisting effects induced by eccentric braking and the like to be considered.
Where restraint capacity is provided by several individual restraints, the load distribution to
each should take account of the effect of construction tolerances of the restraint system, as
well as the relative stiffnesses of portions of the structure.
In general, braking forces are well distributed between piers and abutments. Even deck
joints can resist some force. Any movement of the superstructure resulting from braking
forces is limited by the available travel of the deck joints, and once this is exhausted the
force is resisted by passive earth pressure behind the abutment. It would be rare for piers or
other support elements to be damaged before this happened.

www.standards.org.au  Standards Australia


AS 5100.2 Supp 1—2007 18
Accessed by Brandon & Associates refreshed 170613 on 23 Oct 2019 [OBSOLESCENT] (Document currency not guaranteed when printed)

FIGURE C6.8.2 BRAKING FORCES (F BS) AND (F BM) VERSUS


LENGTH OF STRUCTURE

C6.9 Fatigue load effects


The development of the Standard has provided the opportunity to upgrade the fatigue
criteria of ABDC-1992 (HB 77.2—1996) (Ref. 1) taking into consideration the growth in
heavy vehicle traffic numbers and mass on important freight routes. See Roberts and
Heywood (2004) (Ref. 13), and Grundy and Boully (2004) (Ref. 14).
The fatigue loading has been developed to be consistent with the SM1600 road traffic
design live loading. It is based on the same family of future vehicles and thus is unique to
Australia.
Fatigue is the process of cumulative local structural deterioration of a component or
connection when subjected to repeated, fluctuating stresses. In bridges, it is normally
caused by the repeated loading of a structure by moving vehicles and occurs over an
extended period of time.
This Commentary provides information on fatigue loading. For information on the material
resistance due to the fatigue loading, see AS 5100.5 Supp 1 and AS 5100.6 Supp 1.
Designers should note the importance of producing effective detailing to minimize damage.
In developing the fatigue design criteria, a fatigue life of 75 years has been assumed. This
approach is compatible with a 100 year design life on the basis that—
(a) bridges will be inspected regularly and that intervention will occur when fatigue
damage is detected. This process should ensure that fatigue damage is controlled; and
(b) the uncertainty associated with the prediction of the fatigue life is such that only a
relatively low percentage of bridges for which the theoretical design fatigue life has
been reached actually exhibit fatigue damage.
In developing the fatigue design criteria, actual truck masses and configurations were used.
The data was obtained from weigh-in-motion data for four principal route types and was
also extrapolated into the future, using historical growth rates. The fatigue loading by actual
trucks was then calibrated against the SM1600 and M1600 moving traffic load (without
accompanying UDL) to remove the requirement to introduce further design vehicle loads.
This approach has removed the need for the designer to determine the number and
amplitude of the stress cycles due to the passage of actual trucks. In reality, the passage of a
single vehicle over a bridge produces a number of cycles of stress of varying amplitude.

 Standards Australia www.standards.org.au


19 AS 5100.2 Supp 1—2007

The cumulative fatigue loading of this stress history is converted to the equivalent loading
*
of one cycle of amplitude f equiv ( )
shown in Figure C6.9(A). This is the fatigue stress cycle
specified in the Standard.
Accessed by Brandon & Associates refreshed 170613 on 23 Oct 2019 [OBSOLESCENT] (Document currency not guaranteed when printed)

FIGURE C6.9(A) FATIGUE STRESS CYCLE DUE TO A SINGLE PASSAGE


OF VEHICLE OVER A BRIDGE

Using Miner’s Rule, the design cumulative fatigue loading may be expressed as follows:

( )
Σni f i*
m
( *
= n equiv f equiv )
m
. . . C6.9

where
m = relevant exponent related to fatigue design life (see AS 5100.6)
n equiv = 1
The design cumulative loading caused by the passage of the assumed mix, mass and number
of actual vehicles is expressed as an equivalent number of cycles of 0.70 × (A160 axle
load)(1 + α), in the case of very short spans or 0.70 × (M1600 moving traffic load without
UDL)(1 + α) for other spans.
The maximum stress range due to the passage of the A160 axle load or the M1600 moving
traffic load without UDL is ( f A160 ) m or ( f M1600 ) m , respectively.
Thus the design cumulative fatigue loading can be expressed as follows:

(*
n A160 × f 0.7A160 ) m
; or
*
(
n M1600 × f 0.7M1600 ) m

where
n A160 = number of fatigue stress cycles for the A160 axle load
n M1600 = number of fatigue stress cycles for the M1600 moving traffic load

(f *
)
0.7A160
m = equivalent cumulative
load)(1 + α)
fatigue function for 0.70 × (A160 axle

(f *
0.7M1600
m
) = equivalent cumulative fatigue function for 0.70 × (M1600 moving
traffic load without UDL)(1 + α)

www.standards.org.au  Standards Australia


AS 5100.2 Supp 1—2007 20

The design number of fatigue cycles is applicable to simply supported, continuous and
cantilever spans.
In determining the maximum stress range in the component under consideration, only 70%
of the effect of the A160 axle load or the M1600 moving traffic load is used. This reduction
has been introduced to take the following into consideration:
(i) The actual stresses in a component are generally less than the theoretically calculated
values because of alternative load paths (such as bridge barriers) and the magnitude
of actual components in comparison with line elements used to represent them in
analysis.
(ii) The actual lateral position of heavy vehicles varies and does not generally coincide
with the critical lateral position.
In general terms, the mass of heavy vehicles has been assumed to increase from current
values to that equivalent to SM1600 vehicles over a period of about 50 years. Emphasis has
been placed on the average mass of vehicles rather than maximum allowable mass as the
Accessed by Brandon & Associates refreshed 170613 on 23 Oct 2019 [OBSOLESCENT] (Document currency not guaranteed when printed)

percentage of vehicles loaded to maximum allowable mass may be expected to decrease as


the allowable mass is increased over time and vehicle volume tends to control. The
compound rate of increase of freight is assumed to be about 5% on major interstate routes
down to only about 1% on lesser routes. A saturation volume of between 4000 and 4500
heavy vehicles per lane per day has also been assumed.
For the determination of the number of fatigue stress cycles, a route factor has been
introduced to take into consideration the average mass of heavy freight vehicles on different
types of routes. Vehicles travelling on principal interstate freeways and highways are
generally maximum sized legal vehicles, fully laden and include a high proportion of
medium and long combination vehicles. Vehicles travelling on urban freeways are more
likely to be not fully loaded, with many empty. Vehicles on other rural roads are less likely
to be maximum sized than those on principal interstate routes. Vehicles on urban roads
other than freeways are likely to be shorter and lighter than heavy vehicles on other types of
roads. Considerable use has been made of weigh-in-motion data in determining the different
route factors.
General traffic data has been used to determine the presence of heavy vehicles in different
lanes of multi-lane roads. This data indicated that for rural roads more than 95% of all
heavy vehicles travelled in the slow lane. For urban roads, particularly for freeways with
frequent on and off ramps, a maximum of about 65% of heavy vehicles travelled in a single
lane. This data has been used in specifying the number of heavy vehicles per day per lane to
be used based on the total number of heavy vehicles travelling in that direction.
Guidance on the number of trucks per lane should be sought from the authority.
An indication may be obtained from weigh-in-motion data for Austroads Vehicle
Classifications Heavy Vehicle Classes 3 to 12 (see Figure C6.9(B)). Where weigh-in-
motion data is not available, it may be possible to estimate truck volumes as a proportion of
AADT. Some rationalization of raw data is necessary to ensure a consistent standard for the
design and assessment of bridges on similar parts of the road network.
The following are the typical number of heavy trucks per day in each direction for selected
routes in 2002:
(A) Hume Freeway, 50 km north of Melbourne........................................................ 1500.
(B) Western Highway, near Stawell........................................................................... 500.
(C) Western Ring Road, urban Melbourne ............................................................... 2400.
(D) Monash Highway, urban Melbourne .................................................................. 2100.
(E) Stud Road, near Dandenong, urban...................................................................... 820.

 Standards Australia www.standards.org.au


21 AS 5100.2 Supp 1—2007

(F) Sydney, Newcastle Freeway, 35 km north of Sydney ......................................... 1040.


(G) Pacific Highway, 50 km south of Tweed Heads ................................................... 350.
(H) New England Highway, Maitland........................................................................ 500.
(I) Newell Highway, Dubbo, Forbes, Jerilderie......................................................... 300.
For further information, see Refs 8 to 14.
Accessed by Brandon & Associates refreshed 170613 on 23 Oct 2019 [OBSOLESCENT] (Document currency not guaranteed when printed)

www.standards.org.au  Standards Australia


Accessed by Brandon & Associates refreshed 170613 on 23 Oct 2019 [OBSOLESCENT] (Document currency not guaranteed when printed)

Level 1 Level 2 Level 3


Length Axles and axle Austroads classification
Vehicle type
(indicative) groups
Type Axles Groups Typical description Class Parameters Typical configuration

 Standards Australia
Light vehicles
AS 5100.2 Supp 1—2007

Short Short
(up to 5.5 m) (sedan, wagon, 4WD, utility,
1 or 2 1 d 1 ≤ 3.2 m and axles = 2
light van, bicycle, motorcycle
and the like)
Groups 3
Short-towing
3, 4 d 1 ≥ 2.1 m, d1 ≤ 3.2 m,
3 (trailer, caravan, boat and the 2
or 5 d 2 ≥ 2.1 m and axles = 3, 4
like)
or 5
Heavy vehicles

Medium
(5.5 m to 2 2 Two axle truck or bus 3 d 1 > 3.2 and axles = 2
14.5 m)
22

3 2 Three axle truck or bus 4 axles = 3 and groups = 2

>3 2 Four axle truck 5 axles > 3 and groups = 2

Three axle articulated


d 1 > 3.2 m, axles = 3
3 3 (three axle articulated vehicle 6
and groups = 3
or rigid vehicle and trailer)
Four axle articulated d 2 < 2.1 m or d1 > 2.1 m
4 >2 (four axle articulated vehicle or 7 or d1 > 3.2 m;
Long rigid vehicle and trailer) axles = 4 and groups > 2
(11.5 m to
19.0 m) Five axle articulated d 2 < 2.1 m or d1 > 2.1 m
5 >2 (five axle articulated vehicle or 8 or d1 > 3.2 m;
rigid vehicle and trailer) axles = 5 and groups > 2
Six axle articulated axles = 6 and groups > 2; or
≥6 >2 (six axle articulated vehicle or 9 axles > 6 and groups = 3
rigid vehicle and trailer

www.standards.org.au
FIGURE C6.9(B) (in part) AUSTROADS HEAVY VEHICLES CLASSES 3 TO 12
Accessed by Brandon & Associates refreshed 170613 on 23 Oct 2019 [OBSOLESCENT] (Document currency not guaranteed when printed)

Level 1 Level 2 Level 3


Length Axles and axle Austroads classification
Vehicle type
(indicative) groups
Type Axles Groups Typical description Class Parameters Typical configuration
B Double
>6 4 (B Double or heavy truck and 10 groups = 4 and axles > 6

www.standards.org.au
Medium trailer)
combination
(17.5 m to Double road train
36.5 m) (double road train or medium groups = 5 or 6
>6 5 or 6 11
articulated vehicle and one and axles > 6
dog trailer (M.A.D))
Large
Triple road train
combination groups > 6
>6 >6 (triple road train or heavy truck 12
(greater than and axles > 6
and three trailers)
33.0 m)
where
group = axle group, where adjacent axles are less than 2.1 m apart
groups = number of axle groups
23

axles = number of axles (maximum axle spacing of 10 m)


d1 = distance between first and second axle
d2 = distance between second and third axle

FIGURE C6.9(B) (in part) AUSTROADS HEAVY VEHICLES CLASSES 3 TO 12

 Standards Australia
AS 5100.2 Supp 1—2007
AS 5100.2 Supp 1—2007 24

C6.10 Load factors


Generally, the loads correspond to the serviceability limit state, i.e., load factor is 1.0. The
exceptions are the braking forces which have been reduced due to their low probability of
occurrence (see Clause C6.8.2).
The load factor for the ultimate limit state live load has been set at 1.8. This is a 10%
reduction from the 2.0 used in ABDC-1992 (HB 77.2—1996) (Ref. 1). The reduction is
consistent with international trends, the fact that as legal axle loads increase the proportion
of vehicles that can overload decreases, i.e., their weight is limited by the volume of the
vehicle rather than limits on axle mass and the amount that they can overload is limited.
For centrifugal and braking forces to exist, vehicles should be on the structure.
Consequently, there are co-existing horizontal and vertical loads. This is recognized in the
Standard and presented as a combination that needs to be considered. For example—
 j

1.8  Fc and ∑ ALF i × M 1600 i × (1 + α )
 i =1 
Accessed by Brandon & Associates refreshed 170613 on 23 Oct 2019 [OBSOLESCENT] (Document currency not guaranteed when printed)

requires the centrifugal forces (F c) to be combined with the vertical forces associated with
the M1600 moving traffic load. Note that the centrifugal force relates to the acceleration of
the mass and is thus independent of dynamic effects. This is combined with the vertical
effects of the M1600 moving traffic load including the allowance for dynamics.
C6.11 Deflection
The specification of limits on live load deflections and span to depth ratios have
traditionally been used in road and railway bridge design codes to attempt to control
vibration, prevent fatigue, limit stresses in secondary members and allow for dynamic
loading. Although this Standard provides specific clauses aimed at addressing these
serviceability limit states, it has also retained limitations on live load deflections.
Relaxed deflection limits have been introduced in this version of the Standard as part of the
transition to designing bridges for SM1600 loading and the uncertainty about appropriate
serviceability limits for controlling vibration.
For bridges with walkways, the criteria for limiting vibration in Clause 12 of the Standard
will generally control.
For other bridges, the magnitude of allowable deflection limits have been increased from
1/800 of the span or 1/400 of the cantilever projection in ABDC-1992 (HB 77.2—1996)
(Ref. 1) to 1/600 and 1/300 respectively and the method of calculation has also been made
less conservative, as per AASHTO LRFD (2004) (Ref. 15).
These limits are similar to the values that have been used for railway bridges for many
years and reflect the magnitude of the SM1600 loading. They are also representative of the
magnitude of live load deflections in typical prestressed concrete bridges designed for
SM1600. These deflection limits, together with the fatigue criteria specified in Clause 6.9
of the Standard, are likely to prove to be controlling serviceability limit states for steel
beam bridges, particularly when high strength steel is used.
C6.12 Distribution of load traffic loads through fill
Figure C6.12 illustrates the distribution of wheel loads from road traffic through fill to the
top of buried structures as specified in the Clause and covers individual and overlapping
conditions.
The load distribution is based on that contained within AS 1597.2, making its application
more general to all types of buried structures beneath roadways, as generally recommended
by Standards Australia Committee CE-025.

 Standards Australia www.standards.org.au


25 AS 5100.2 Supp 1—2007
Accessed by Brandon & Associates refreshed 170613 on 23 Oct 2019 [OBSOLESCENT] (Document currency not guaranteed when printed)

FIGURE C6.12 DISTRIBUTION OF WHEEL LOADS THROUGH FILL

More refined methods of analysis may be used to derive more accurate distributions, if
approved by the relevant authority.
Whilst the Clause specifies the distribution through fill of the SM1600 design loads, the
requirements of the Clause also apply to heavy load platforms.
The depth of fill includes the pavement and should be measured to the top of the finished
wearing surface.

www.standards.org.au  Standards Australia


AS 5100.2 Supp 1—2007 26

The uniform load components of 6 kN/m (1.875 kPa) of the M1600 moving traffic load and
24 kN/m (7.5 kPa) of the S1600 stationary traffic load are not dispersed longitudinally, but
are dispersed transversely as for wheel loads, as the structure width may be wider than the
roadway width.
The width of the distribution cannot be wider than the structure supporting the roadway.
For buried structures, the dynamic load allowance should be applied to the road traffic
loads as specified in Clause 6.7.3 of the Standard.

C7 PEDESTRIAN AND BICYCLE-PATH LOAD


C7.1 General
It is unlikely that full maximum load will occur over a large walkway area, except in
extreme crowds, e.g., major sporting events.
For walkways attached to and supported by the main bridge superstructure members, the
design pedestrian load to be carried by these members decreases as the area of walkway
increases. Individual components supporting smaller portions of the walkway area are to be
Accessed by Brandon & Associates refreshed 170613 on 23 Oct 2019 [OBSOLESCENT] (Document currency not guaranteed when printed)

designed for the high crowd loads possible over these small areas.
For pedestrian bridges and independent walkways, the structural members are to be
designed to carry the higher pedestrian loads, especially if the location of the bridge is such
that crowd loads are likely.
C7.2 Service live load on walkways
The additional load for walkways and service platforms is to provide for the stacking of
materials during construction and servicing.
C7.3 Load factors
The Clause specifies a load factor of 1.8 for pedestrian load, which is greater than the value
of 1.5 specified in AS/NZS 1170.1, Structural design actions—Permanent, imposed and
other actions. The value of 1.8 is consistent with providing an acceptable probability that
the design load will not be exceeded during the 100 year design life taking into
consideration unexpected crowd loading and the like. It also ensures an increased level of
robustness which is important for road and bridge related structures that may be exposed to
accidental collision loads.

C8 RAILWAY TRAFFIC
C8.1 General
The railway traffic loads specified in the Standard, including the various factors, have been
determined to cover the effects of—
(a) the maximum expected loading to be applied; and
(b) a large number of load repetitions of the operating loads which defines the fatigue
loading spectra.
Live loads other than the specified 300LA may be used at the discretion of the railway
authority, but this should be a factored version of the 300LA loading. This should not be
done without good reason, and only after noting that lines which are currently only lightly
loaded may have their loading profiles changed at a future time.
C8.2 300LA railway traffic load
The 300LA loading combines the two alternatives of the 300-A-12 loading from
ABDC-1992 (HB 77.2—1996) (Ref. 1), by adding the 360 kN single axle load 2 m in front
of the vehicle loading. This is to simulate six axle coupled locomotives and better represent
their loading in the 15 to 22 m span range. This combination produces moment and shear
envelopes more nearly proportional to that of existing trains over the whole range of spans
[see Marcer (2002) (Ref. 16)].
 Standards Australia www.standards.org.au
27 AS 5100.2 Supp 1—2007

For bridges carrying special purpose rail loading, e.g., light rail and cane rail, the loading
will depend on the actual vehicle configurations.
C8.3 Multiple track factor for railway bridges
The load modification factors given in Table 8.3 are to take into account the reduced
probability that extreme loads will occur simultaneously on all tracks of multiple-track
railway bridges.
In assessing the worst effects that are to be considered in each member, all possible
combinations of simultaneous track loading should be considered, and if the number of
tracks being loaded is greater than 2, the loading may be reduced using the relevant load
modification factor.
C8.4 Dynamic load allowance
C8.4.1 General
The dynamic load allowances for railway live loads in the Australian Bridge Design Code
were based upon the UIC Code Leaflet 776-1 R (1994) (Ref. 17), and assumed perfect
Accessed by Brandon & Associates refreshed 170613 on 23 Oct 2019 [OBSOLESCENT] (Document currency not guaranteed when printed)

wheels, i.e., worn wheels were not considered. Strain gauging of many railway bridges in
NSW and comparison with published strain gauging from USA and Canada has resulted in
modification to the dynamic load allowance to allow for the lower standard of track and
worn wheels in these countries compared to Europe (Refs 16, 18 and 19).
On the basis of the above considerations, the dynamic load allowance for railway live load
is now defined as the dynamic load effect occurring with the member under consideration,
with the type of wheel defect maximizing dynamic load effect for that member, with a
wheel defect at the condemning limit. A wheel defect at the condemning limit is defined as
the limit beyond which a vehicle is not permitted to run. The maximum dynamic load is
usually recorded with a train speed in the range 80 to 100 km/h or higher and with
significantly worn disc-braked wheels, with circumferential defects. These produce higher
rates of strain than out of round wheels and wheel flats. The highest rates of strain
maximize fatigue damage, the possibility of crack propagation and the possibility of brittle
fracture.
Dynamically balanced steam locomotives can be considered to comply with the given
dynamic load allowances. Hammer blow is only likely to exceed the given dynamic load
allowance for non-dynamically balanced steam locomotives at speed. Where there is no
evidence of dynamic balancing, a higher dynamic load allowance may be applied, such as
the American Railway Engineering Association, AREA (1969) (Ref. 20), for steam
locomotives with hammer blow; provided this gives a higher value.
Locomotive 3801, which was dynamically balanced from manufacture, has been strain
gauged at Rockdale, Harrow Road Railway Bridge (Ref. 19) to have no worse dynamic load
than 81 class locomotives with good wheels at about 70 km/h.
Members with high L/r ratio are subject to high dynamic load and resonance.
For structures such as trusses where members are acting in tension or compression, the
dynamic load allowance for each member should be based on the load action being resisted,
e.g., truss chord members designed for α for bending movement, truss diagonals designed
for α for shear.
It is also possible to derive relationships between the dynamic load allowance and train
speed, and these relationships also depend on the assumed bridge natural frequencies and
the level of damping. They can be simplified to the following rules, which are conservative,
and apply to the bending moment, shear and reactions. Strain gauging of various captive
trains may give a more accurate relationship for particular bridges.

www.standards.org.au  Standards Australia


AS 5100.2 Supp 1—2007 28

C8.4.2 Characteristic lengths (L α )


Characteristic length is a key parameter in determining the natural period of vibration of a
bridge as well as well as its response to suddenly applied loads.
C8.4.3 Dynamic load allowance for bending effects
The values given in the Standard are based on the best available information for Australian
condition. Further research may be necessary.
C8.4.3.1 Ballasted deck spans
The values given in Table 8.4.3.1 of the Standard are those given in ABDC-1992
(HB 77.2—1996) (Ref. 1).
C8.4.3.2 Open deck spans and spans with direct fixation
The values given in Table 8.4.3.2 are those from extensive strain gauging of transom top
spans in NSW and comparison with CP Engineering research and AAR research (Ref. 18)
with wheels machined to condemning limits, one of which is also the NSW condemning
Accessed by Brandon & Associates refreshed 170613 on 23 Oct 2019 [OBSOLESCENT] (Document currency not guaranteed when printed)

limit.
C8.4.4 Application
(No Commentary)
C8.4.5 Dynamic load allowance for other load effects
The derivation of the dynamic load allowance for shear, torsion and reactions as 2/3 of the
value of the bending moment is from UIC-776-1 R (1994) (Ref. 17).
C8.4.6 Dynamic load reversal
Bridges continue to vibrate after the passing of vehicles and this can reduce the minimum
reaction or even cause uplift. By neglecting the attenuation of the dynamic response, it
follows that the magnitude of this effect is approximately the same as the dynamic load
allowance.
This phenomenon is suspected as a contributory cause of apparent low frictional resistance
of bearings, the walking of bearings, fatigue failures of metal components of deck joints
and also of flexural cracks on top of light prestressed concrete beams.
From strain gauging of railway bridges, the dynamic load allowance for the worst wheels at
speed is calculated by comparing the mean of the maximum oscillation with the maximum.
This may be checked against a similar train at crawl or slow speed.
C8.4.7 Application to dedicated lines and traffic
For standard track, the dynamic load allowance can be taken to be constant for speeds
above 80 km/h, and to vary linearly from 0.00 for a speed of zero to the full value at
80 km/h.
C8.5 Distribution of railway traffic load
C8.5.1 General
(No Commentary)
C8.5.2 Open deck steel rail bridges
Strain gauging and second order analysis of timber transoms have shown these assumptions
to be suitably conservative.
C8.5.3 Ballasted deck steel railway bridges
(No Commentary)

 Standards Australia www.standards.org.au


29 AS 5100.2 Supp 1—2007

C8.5.4 Ballasted deck concrete railway bridges


(No Commentary)
C8.5.5 Direct fixation
(No Commentary)
C8.6 Horizontal forces
C8.6.1 Centrifugal forces
Although the Clause specifies that centrifugal force shall not be increased by the dynamic
load allowance, experimental evidence indicates that this effect is significant in terms of the
fatigue damage. AS 5100.6, Bridge design—Steel and composite construction, specifies an
empirical allowance to allow for this effect.
The load resulting from the centrifugal force should always be combined with a vertical
load.
For curved bridges, the load case of full vertical live load without centrifugal forces should
Accessed by Brandon & Associates refreshed 170613 on 23 Oct 2019 [OBSOLESCENT] (Document currency not guaranteed when printed)

also be considered.
C8.6.2 Braking and traction forces
The design braking and traction forces specified in the Standard reflects values for friction
between rail and wheel which have been measured to be up to the order of 0.3 to 0.4.
Traction forces apply to the locomotives whereas braking forces can apply to all axles. The
values for longitudinal force given in Table 8.6.2 of the Standard have been simplified, and
can be taken to apply to both braking and traction.
The transmission of longitudinal forces into the structure depends on the system of rail
support. If the rails are non-continuous then the longitudinal forces should be resisted by
the bridge substructure, but if the rails are continuous then some of the longitudinal forces
will be transmitted to the approach embankments via the rails. For bridges with continuous
tracks, the forces can be considered to be distributed by the rails to the various parts of the
structure depending on their stiffnesses.
Continuous rails for the purposes of the Clause are those rails connected by welded joints,
and which extend continuously from embankment to embankment, with an effectively
anchored length of at least 20 m past the ends of the bridge.
Where restraint capacity is provided by several individual restraints, the load distribution to
each should take account of the effect of construction tolerances of the restraint system, as
well as the relative stiffness of portions of the structure.
An analysis which models the stiffness of all bridge components, including the continuous
rails, should be considered for large or unusual structures or where there is partial fixity due
to points or crossings near the ends of the bridge.
C8.6.3 Nosing loads
Although the Clause specifies that nosing load shall not be increased by the dynamic load
allowance, experimental evidence indicates that this effect is significant in terms of the
fatigue damage. AS 5100.6 specifies an empirical allowance to allow for this effect.
The requirement to design for nosing load is intended to cover a variety of requirements,
some of which are difficult to quantify. These include the—
(a) lateral loadings created by the passage of trams;
(b) requirement that railway bridges and their individual members such as steel stringers,
have a certain minimum lateral stiffness;
(c) effect of trackwork on bridges;

www.standards.org.au  Standards Australia


AS 5100.2 Supp 1—2007 30

(d) effect of rail breaks on bridges; and


(e) effect of minor impacts on the superstructure.
The nosing load may not always cover the effect of the above, and the designer should be
aware that in extreme situations, the forces created by the interaction between the track and
the structure may require special consideration.
The effects of trackwork on bridges may include the following:
(i) For curved track, a horizontal lateral serviceability load (H TR) on each track, acting
towards the centre of the curve and applied at track level equal to—
H TR = T/R . . . C8.6.3
where
T = rail tension at minimum temperature, in kilonewtons
R = radius of curvature of track, in metres
Accessed by Brandon & Associates refreshed 170613 on 23 Oct 2019 [OBSOLESCENT] (Document currency not guaranteed when printed)

(ii) A longitudinal force applied at track level as a result of a break in a continuous rail,
usually in association with cold temperature conditions.
In many cases, the effect of the nosing load will be less critical than the minimum ultimate
lateral restraint capacity, but the effects on bearings and fixings should be considered.
C8.7 Fatigue load
C8.7.1 Fatigue design traffic load
Fatigue loadings for railway bridges fall into two distinct categories—
(a) fatigue design loadings (i.e., total loadings) which should not be exceeded (see
AS 5100.5, Bridge design—Concrete); and
(b) fatigue design stress range and the associated number of stress cycles, which cause
steel fatigue at the tip of micro cracks (see AS 5100.5, Bridge design—Concrete and
AS 5100.6, Bridge design—Steel and composite construction). This stress range and
the number of cycles are assessed on the basis of the fatigue design traffic loadings,
i.e., without permanent effects.
The specified load against which these limits are to be checked is the fatigue design traffic
load as specified in the Clause plus permanent effects. For concrete structures, there is a
need to limit the maximum concrete stresses under repeated loads, and this standard
requirement is not related to the number of loading cycles.
C8.7.2 Fatigue design stress range (f *)
For steel structures and for the steel in concrete structures, there is a need to check the
effects of repeated loads. For these checks, the Clause specifies the fatigue design stress
range (f *) and Clause 8.7.3 specifies the associated effective number of stress cycles (n).
The Clause is based on a given constant fatigue load, but if a variable fatigue load can be
reasonably defined, the effect of the variable load may be assessed using the principles set
out in AS 5100.6. The length of the loading to be taken in determining the design stress
range is that which creates the largest structural actions in the member under consideration.
This is based on the assumption that the largest stress is the one that dominates in the
fatigue situation, and is reasonable because the fatigue effect of one stress cycle is
proportional to the cube of its stress range.
The fatigue design stress range (f* ) allows for the maximum permissible design train, and
the fact that, typically, trains are considerably lighter than the design load would suggest is
allowed for by reducing the effective number of stress cycles (n) (see Clause 8.7.4 of the
Standard).

 Standards Australia www.standards.org.au


31 AS 5100.2 Supp 1—2007

The fatigue design stress range (f *) is the appropriate stress range for assessing whether the
constant amplitude fatigue limit has been exceeded since it is the maximum stress range
that should occur.
C8.7.3 Effective number of stress cycles (n)
For simplicity of computation, the actual number of cycles of variable amplitude in this
Commentary is replaced by an equivalent number of constant amplitude stress cycles of the
fatigue design stress range. This range is the maximum design amplitude of stress.
The number of cycles to be considered also depends on the characteristic length of the
loading, because this causes short members (such as those subjected to one load cycle per
axle or bogie) to have a larger number of effective stress cycles than main beams which
may only receive one load cycle per train.
The value given in the Clause is a conservative approximation and assumes that the leading
portion of the load (locomotives) does not significantly add to the fatigue damage.
A more precise estimation of n T may be calculated using estimated values of stresses from the
Accessed by Brandon & Associates refreshed 170613 on 23 Oct 2019 [OBSOLESCENT] (Document currency not guaranteed when printed)

load anticipated to be the predominant train.


For a train consisting of n L locomotives of six axles and nw wagons of four axles, a more
accurate (but still conservative) estimation of n T is given in Table C8.7.3.

TABLE C8.7.3
VALUES OF nT
Lf nT

  fL 
3
 f L max. 
3   
3
 f Wmax. 
3

L f < 9.0 nT = nL 1 + 5 1 − min.     + n 1 + 3 1 − f Wmin.    


    f*   w     f*  
 f L max.   T   f Wmax.   T 
   

 3  f 3  3  f 3
fL    
L f > 9.0 nT = nL 1 − min.  + 1  L max.  +n 1 − f Wmin.  + 1  Wmax. 
    f*  w     f* 
f L max. f Wmax.
    T      T 

where
f L min. = minimum stress under the locomotives (i.e., whilst train on span)
f L max. = maximum stress under the locomotives
f Wmin. = minimum stress under the wagons (i.e., whilst train on span)
f Wmax. = maximum stress under the wagons
f T* (
= max. f L* , f W
*
)
f L* = maximum stress range under locomotives
*
fW = maximum stress range under wagons

C8.7.4 Track category for fatigue load


The base number of loading cycles (C T ) has been based on American work, modified to suit
Australian conditions. The American work found that the actual stress range experienced by
bridge elements is considerably lower than would occur under the design loading. To
compensate for this, the effective number of stress cycles that is considered in design has
been made less than the actual number of load cycles to which a bridge is subjected.
Stresses measured in real bridges in the USA have been found to follow a Rayleigh
probability distribution function, see Fisher (1977) (Ref. 21) correspond to Cooper loadings
in the range of E40 to E55, 50% to 69% of the normal design loading of Cooper E80. This
American data has been used to modify the effective number of stress cycles by
conservatively assuming that actual loads are at the highest end of the range of the
American range, and that the root mean square value of the actual load is 0.72 times the
design load.

www.standards.org.au  Standards Australia


AS 5100.2 Supp 1—2007 32

The railway load for checking repeated loads is based on a track which is used for a variety
of trains at near-saturation traffic levels. The basic design case is the AREA (Ref. 20)
assumption of 60 trains per day per track and 60 coupled units per train. This is equivalent
to the UIC (Ref. 17) assumption of 27 million tonnes of traffic per year per track. If the
design life is assumed to be 100 years, the value of n can be derived as follows:
Effective root mean square load is assumed to be—
M(RMS) = 0.65αM(Design) (as described in this Clause) . . . C8.7.4(1)
where
α = factor relating observed maximum stress range to computed maximum stress
range based upon design load and dynamic load allowance
The value of α is conservatively assumed to be 0.9 for all spans, and hence—
α = 0.9 × 0.72 × f * . . . C8.7.4(2)
= 0.648f * . . . C8.7.4(3)
Accessed by Brandon & Associates refreshed 170613 on 23 Oct 2019 [OBSOLESCENT] (Document currency not guaranteed when printed)

The equivalent number of cycles (n) of f * over 100 years at 60 trains per day is given as
follows:
n = 365 × 100 × 60(0.648) 3n T . . . C8.7.4(4)
= 595,894n T . . . C8.7.4(5)
n T is the equivalent number of stress cycles of maximum amplitude per train as given in
Table 8.7.3 of the Standard. n T is dependent on the span or, more accurately, the base length
of the influence line, since it is only loads located within that length that are effective in
creating the stress that is being considered. The shorter the loaded length considered, the
greater will be the variation in the loading, and to use a full span length would therefore be
non-conservative.
The constant has been rounded up to 600 000 and Equation C8.7.4(4) has been found to
give a good fit to the AREA (Ref. 20) tabulations.
In reviewing these values for Australian practice, the constant has been varied depending on
the different class of line, and three types of line, with different constants (C T) have been
defined as follows:
(a) Heavy haul lines Heavy haul lines are lines such as those carrying minerals to ports,
where the trains are block trains (all fully loaded), and the line may be carrying
50 million tonnes of freight or even more each year. In an extreme case, the constant
(C T ) could be as high as 1 200 000 calculated as follows:
α = 0.9 (due to the consistency of load)
RMS load = 0.9 × 0.9
= 0.81
number of trains = 2 190 000
(60 trains a day for 100 years)
CT = (0.81) 3 × 2 190 000
= 1 163 856 (say 1 200 000)
This is an extreme case, and for such a case, it is unlikely that the life of the mine
would realistically be 100 years, and so 600 000 has been adopted for C T (as derived
for the American case) but nevertheless, in extreme cases, designers may need to
increase the number above 600 000.

 Standards Australia www.standards.org.au


33 AS 5100.2 Supp 1—2007

(b) Main line freight Main line freight lines are considered to be those carrying regular
freight between state capitals and the like. For these lines, the constant (C T ) has been
derived as follows:
α = 0.6 (taken from the American experience)
RMS load = 0.6 × 0.9
= 0.54
number of trains = 730 000
(20 trains a day for 100 years)
CT = (0.54) 3 × 730 000
= 114 949 (say 100 000)
(c) General freight General freight lines are those carrying freight but not on high
density routes, and the category also covers less heavily trafficked grain lines and the
like. For these lines, the constant (C T ) has been derived as follows:
Accessed by Brandon & Associates refreshed 170613 on 23 Oct 2019 [OBSOLESCENT] (Document currency not guaranteed when printed)

α = 0.65 (for general freight)


RMS load = 0.65 × 0.9
= 0.585
number of trains = 52 000
(10 trains a week for 100 years)
CT = (0.585)3 × 52 000
= 10 410 (say 10 000)
(d) Grain lines For grain lines, the constant (C T ) has been derived as follows:
α = 0.9 (assuming heavier wagons, all filled to capacity)
RMS load = 0.9 × 0.9
= 0.81
number of trains = 10 000
(100 trains a year for 100 years)
CT = (0.81) 3 × 10 000
= 5314 (less than the adopted value of 10 000)
C8.7.5 Multiple track bridges
The large variation in the load effects applying to the fatigue of elements in multiple track
bridges makes it impossible to be definitive as to the rules to be applied. It is noted that the
likelihood of repeated load effects in an element from coincident loads on more than one
track is low, and that few elements in a bridge will receive major loads from more than one
track.
For assessing the fatigue design load, the load from the second and subsequent tracks may
be reduced to 80% of the design load, without the dynamic loading allowance. For the
fatigue stress range and the associated number of stress cycles, a rain flow analysis of
cumulative damage may be carried out, and this would normally result in the effects of the
load from the other tracks being very low.
Designers will need to use their own judgement, based on current knowledge, in regard to
the estimation of the number of load events, since the traffic on a heavily used railway
carrying a bulk commodity is quite different to the traffic on a line carrying general freight.

www.standards.org.au  Standards Australia


AS 5100.2 Supp 1—2007 34

The cumulative damage calculation should be undertaken to allow for the fact that the
fatigue damage caused by a stress cycle is proportional to the cube of its stress range.
C8.8 Load factors
(No Commentary)
C8.9 Deflection limits
(No Commentary)

C9 MINIMUM LATERAL RESTRAINT CAPACITY


The provision of the Clause is aimed at preventing lateral displacement of the
superstructure or piers by accidental impact loads and minor earthquakes. Such
displacement may cause damage out of all proportion to the original accidental forces.
Only the ultimate capacity needs to be considered in design, as serviceability loads are not
critical for such restraints.
Accessed by Brandon & Associates refreshed 170613 on 23 Oct 2019 [OBSOLESCENT] (Document currency not guaranteed when printed)

Where a bridge superstructure is such that fairly large lateral displacements would not
cause significant damage, the designer may wish to waive this provision. However, it
remains a principle of good design practice to provide a positive connection or restraint in
preference to relying on gravity and friction.
Where impact on the superstructure from passing road or rail vehicles, or vessels is
possible, the designer should consider the stability of the outer girders and the advantages
of diaphragms placed in likely impact positions to stabilize and protect girders. If a
roadway underneath is on a gradient or an over-bridge has cross-fall, impact on interior
girders is also possible.
In addition, lateral impact may be associated with uplift, e.g., high loads jammed under
bridges, hence, the restraint system should be capable of resisting reasonable upward forces
or displacements of the superstructure. This may be a particular problem with lighter
superstructures, e.g., steel beams or short, simply supported spans.

C10 COLLISION LOADS


C10.1 General
While slender bridge supports may be aesthetically pleasing, they are susceptible to
collapse under forces due to collision. Where collision on supports is possible, the designer
should design for redundancy or protect the supports with rigid barriers, or design the
supports to withstand collisions.
C10.2 Collision load from road traffic
The load of 2000 kN, although increased from the value in ABDC-1992
(HB 77.2—1996) (Ref. 1), still does not necessarily represent the maximum force generated
in a collision by a legally loaded vehicle. The maximum force depends on the flexibility of
the support and its mass, and the crushing characteristics and velocity of the colliding
vehicles. If the support is adjacent to a high speed roadway (greater than 60 km/h), higher
collision forces should be investigated.
For supports on a skew or adjacent to curved roadways, other angles of possible collision
should be considered. If the foundations, or required support profile, are such that they
cannot safely sustain the design collision force, then a rigid support protection barrier
should be installed. Flexible steel W-beam guardrails are not capable of stopping a heavy
vehicle; hence, concrete safety barriers are recommended (see Clause 11 of the Standard).
C10.3 Loads on protection beams
For new bridges, preference should be given to heavy collision resistant superstructures.

 Standards Australia www.standards.org.au


35 AS 5100.2 Supp 1—2007

For existing bridges of lightweight construction, it is advisable to fit a protective beam or


gantry in front of the structure to absorb possible collision forces. The beam should be
placed below the level of the underside of the bridge deck.
It is acceptable for protective beams to undergo considerable deformation under the
specified ultimate load. The distance between the protective beam and the side of the bridge
should be a minimum of 500 mm for deformable beams, and a minimum of 100 mm for
rigid beams. The beams should be supported in such a way that they will continue to be
held in place, and cannot fall onto the roadway due to a major collision, including a
collision from a vehicle coming out from under the bridge and pushing the beam away from
the bridge.
The infrastructure owner may specify reduced loadings consistent with the design of
previous protection beams that have been shown to perform satisfactorily.
C10.4 Collision load from rail traffic
C10.4.1 General
Accessed by Brandon & Associates refreshed 170613 on 23 Oct 2019 [OBSOLESCENT] (Document currency not guaranteed when printed)

The Clause includes ramps and stairways that are part of the structural members spanning
across the track, but not separate ramps and stairways adjacent to the track.
The location of columns and the likelihood of train collision should be considered when
designing new bridges over railways, and additional measures may be necessary to
minimize the likelihood of collisions, or to reduce the possible effect of collisions. If
necessary, the minimum design forces specified in the Standard may need to be increased.
The likelihood of trains becoming derailed is dependent on the track, and the following
track components and track geometry configurations are the factors and components most
likely to cause derailments:
(a) Catch points and derailers, designed to cause a derailment when a train fails to stop in
certain circumstances.
(b) Slips (track component).
(c) Scissors crossovers.
(d) Facing turnout.
(e) Diamonds.
(f) Trailing turnout.
(g) Curved track, particularly on descents and especially at the bottom of long descent.
(h) Straight track on a descent and especially at the bottom of long descent.
Items (a) to (h) are arranged with the factors most likely to cause derailments being given
first. These items are critical for consideration in a risk analysis to satisfy the Standard.
In a major derailment on curved track, the train will travel along the tangent to the track at
that point. In a minor derailment of one bogie, the train will travel parallel to the track.
Where the location of a bridge is such that its columns have a high risk of being hit by a
derailed train, they should be detailed to ensure that they will behave in a ductile way under
the effect of a major collision, and their detailing should be such that repairs can be made
readily and with minimum disruption to traffic. Consideration may also be given to
designing the superstructure to be capable of supporting the deck dead load plus 20% of the
live load at the ultimate limit state with one column removed.

www.standards.org.au  Standards Australia


AS 5100.2 Supp 1—2007 36

C10.4.2 Provision of alternate load path


The provision of bridges and structures above railways having adequate alternative load
paths to allow removal of the vulnerable support during a train derailment or collision, is
far more important than meeting the load requirements of Clause 10.4.3 of the Standard.
This is because trains are involved in incidents frequently above the speeds, specified in
Clause 10.4.3, that the loads allow for. A derailment on a bridge in NSW has occurred at
160 km/h within the last 20 years. Provision of sufficient alternate load paths rather than
supports of heavy construction, has the advantage that when the train hits square on to the
support, as at Granville (Ref. 22), the train will proceed with little interference to the
direction of its progress, as the 160 km/h derailment did.
C10.4.3 Collision loads on support elements
The pier ultimate design loads of 3000 kN parallel to the rail and 1500 kN normal to the rail
are chosen to satisfy most rail authorities’ requirements. They are not necessarily the
maximum forces that would be generated.
The requirements of the Clause are intended to satisfy the conditions of moderate
Accessed by Brandon & Associates refreshed 170613 on 23 Oct 2019 [OBSOLESCENT] (Document currency not guaranteed when printed)

derailments and minor collisions, but no major derailment of a 300LA train with 84 wagons
derailing at slow speed or collision of two trains of this type at shunting speeds. (Some
systems operate with 84 wagons in trains approximating 300LA.) This may also represent a
self-propelled passenger train derailing at moderate speed or a collision of two trains of this
type at slow speed. Where these provisions are likely to be exceeded, the design should
allow for a pier or one or two columns to be removed.
For flyovers, a risk analysis would be desirable to assess whether the load in these clauses
needs to be increased, or other provisions made.
A solid concrete full height pier, 600 mm or more thick and 4 m or more wide, with 1%
vertical steel and with adequate connection to the superstructure should in most cases
satisfy these load conditions.
Where a freight train derails and impacts head on with rigid pier protection at a speed
approaching 80 km/h, sufficient kinetic energy is stored to permit the locomotive to rotate
end over end on to the superstructure; thus, alternative load paths are a much better option
in such a case.
C10.4.4 Bridge and structural components within 10 m of the centre-line of the railway
track
The forces specified in the Clause are representative of those that may be generated in a
train collision between two trains at the speeds specified in Clause 10.4.3, where some train
vehicles become airborne. The possibility of this occurring at platforms is increased due to
the limitation of the colliding train’s lateral movement, by the platform. At stations where
trains are regularly shunted up to one another, the probability of this occurring increases.
C10.4.5 Underground railway, air space developments and similar situations
Because of the consequences and difficulty of rescue, particularly if fire is involved, the
Clause has been included to justify more stringent requirements for these locations.
Where the railway authority’s operational procedures require that train stops or approved
equivalent devices are used for all trains, the loads specified in the Clause may be reduced
to a level not less than that specified in Clause 10.4.4. This reduction may also apply to
general rolling stock travelling at speeds not greater than 20 km/h.
C10.4.6 Other design requirements
(No Commentary)

 Standards Australia www.standards.org.au


37 AS 5100.2 Supp 1—2007

C10.5 Derailment loads


C10.5.1 General
The objective of checking for derailment loadings is to ensure that when a derailment
occurs, the structure will not suffer a disproportionate amount of damage in the
superstructure and the substructure, or as a result of stability problems. In a derailment
where the train leaves the track but does not move laterally more than about 1.3 m, the
major structural components should not suffer more than superficial damage (minor
components need not be considered). In a serious derailment, the train should not cause the
essential structure of the bridge to be damaged beyond economic repair, due to either
structural failure or to instability. The structure should also be able to support a train that is
loading the edge of the deck.
The Standard does not require that derailed trains be retained on a bridge structure, which is
a matter for the infrastructure owner. If the retention of derailed trains on a bridge
superstructure is a design requirement, Clause 10.4.3 of the Standard has to be used.
C10.5.2 Derailment load Case A
Accessed by Brandon & Associates refreshed 170613 on 23 Oct 2019 [OBSOLESCENT] (Document currency not guaranteed when printed)

Derailment load Case A envisages the derailment of locomotives or heavy freight wagons,
with the derailed vehicles remaining in the track area on the bridge deck.
C10.5.3 Derailment load Case B
Derailment load Case B envisages the derailment of locomotives or heavy freight wagons,
with the derailed vehicles leaving the track area without falling off the bridge, but
remaining balanced on its edge.

C11 KERB AND BARRIER DESIGN LOADS AND OTHER REQUIREMENTS FOR
ROAD TRAFFIC BARRIERS
C11.1 Kerb design loads
This design load is based on the NAASRA Bridge Design Specification (1976) (Ref. 23).
C11.2 Barriers
C11.2.1 General
Minor detail changes or improvements may be made to a tested system provided they do not
detract from the performance of the original tested system, subject to the approval of the
authority.
Sound engineering judgement should be used when making any modifications to crash-
tested systems to ensure that these changes do not adversely affect these systems. The
evaluation of the effects of any changes should take into consideration original crash-test
results and any supporting calculations and component testing. In general, the more flexible
the system, the more likely that changes may adversely affect the performance of the
system.
Reference should be made to other Standards and specialist literature (see Refs 23 to 37).
C11.2.2 Traffic barrier design loads
The design loads have been chosen to align with values in the AASHTO-2004 LRFD
Specifications (Ref. 15). This choice ensures that with minor improvements, barrier systems
successfully tested in accordance with the requirements of the relevant levels in the
internationally accepted testing document [NCHRP Report 350 (Ref. 24)] can be adopted.
Each ultimate design load combination is applied to all parts of the road bridge traffic
barrier system, which contains these loads by deforming and absorbing the energy of
impact.

www.standards.org.au  Standards Australia


AS 5100.2 Supp 1—2007 38

The design loads are considerably greater than those quoted in previous NAASRA Bridge
Design Specification (1976) (Ref. 23) and the ABDC-1992 (HB 77.2—1996) (Ref. 1),
which were applied to a section and distributed to a localised area.
In Table 11.2.2 of the Standard, the lengths for L T , L L and LV relate to the length of
significant contact between the vehicle and barrier that has been observed in films of crash
tests. The length of 1.1 m for the regular performance level is the rear axle tyre diameter of
the truck. The length of 2.4 m for the medium performance level, as described in
Appendix A of the Standard, is the length of the prime mover or trailer rear tandem axles,
two 1.05 m diameter tyres, plus 300 mm between them. The weight of the vehicle (F V)
lying on top of the bridge barrier is distributed over the length of the vehicle (L V) in contact
with the barrier.
For detailed guidance on the design of prototype barriers for load testing and the methods
of applying loads to barriers, see AASHTO LRFD Bridge Design Specifications (2004)
(Ref. 15). Sound engineering judgment should be used when proportioning loads to
different components of barrier systems to ensure that the structural and geometric details
of such systems are compatible with the relevant characteristics of all vehicles to be
Accessed by Brandon & Associates refreshed 170613 on 23 Oct 2019 [OBSOLESCENT] (Document currency not guaranteed when printed)

contained for the specified performance level.


C11.2.3 Effective height
The methodology for calculating the effective height of the barrier taking into account the
geometry and stiffness of the different barrier components, the distribution of loads
between longitudinal rails and the centre of gravity of the controlling vehicle, is provided in
the AASHTO LRFD Bridge Design Specifications (2004) (Ref. 15).
The designer should consider the ultimate behaviour of the barrier under impact including
the potential for lateral rotation due to yielding of member sections with the resultant
potential for vehicle ramping. For this purpose, the lateral rotation of a post and rail barrier
should be limited to a maximum of 40° under test vehicle impacts.
The actual height of rigid parapets should be marginally higher than the required effective
height to allow the effective distribution of the test vehicle impact loading to the barrier
elements.
The minimum effective heights of the low, regular and medium performance level barriers
are based upon theoretical assessments supported by overseas testing and in-service
performance, to prevent the relevant vehicles given in Table 10.4 of AS 5100.1 from
overturning these barriers under the specified impact criteria.
The low performance level minimum effective height of 500 mm is based upon the tyre
height of the test vehicles and other light vehicles.
The regular performance level effective height of 800 mm is based on the theoretical
requirement and on satisfactory performance in preventing light test vehicles and the
8 tonne rigid test truck rolling over such barriers, under the specified impact criteria
(Ref. 15).
For commentaries to the minimum effective heights for medium and special performance
level barriers, see Paragraph CA3 of Appendix CA.
C11.2.4 Anchorage
To minimize damage to bridge decks, it is necessary to design progressive strength systems
in which road bridge barriers and their connections fail prior to the supporting structural
elements. To ensure this behaviour, in lieu of a detailed structural analysis the barrier/deck
anchorages should be designed in accordance with the Standard to have a structural strength
greater than the connecting elements at the base of the barrier.

 Standards Australia www.standards.org.au


39 AS 5100.2 Supp 1—2007

C11.2.5 Continuity
Within reasonable economic limits, a road bridge barrier should provide moment, shear, and
tensile continuity throughout its length. Providing these features will require designing end
anchorages and continuity transfer splices and expansion devices for beam-and-post barriers
and continuity transfer devices for open joints in concrete parapets.
In the case of concrete parapet type barriers, continuity enables longitudinal beam actions
as well as vertical cantilever action from the bridge deck to contribute to the ultimate
containment capacity of the section of barrier. Attention needs to be given to the design of
continuous cast in situ concrete parapets to minimize shrinkage cracking.
In the case of steel railing, 75% tensile continuity in the splice section is specified as a
minimum to allow for the reduction in strength due to bolt holes within the continuity
transfer splice.
C11.3 Bridge deck cantilevers
For decks supporting rigid parapets, deck overhang capacity should be not less than the
Accessed by Brandon & Associates refreshed 170613 on 23 Oct 2019 [OBSOLESCENT] (Document currency not guaranteed when printed)

parapet base anchorage capacity, otherwise the failure mechanism may not be restricted to
the parapet alone. For such cases, permanent damage can occur to the cantilever, following
a heavy vehicle-parapet impact.
When impacted, the post of a post and rail barrier imposes large concentrated loads on the
bridge deck at the point where the post is attached to the deck. Concrete slabs or decks can
fail in punching shear resulting from the force in the compression flange of the post.
Adequate thickness or edge distance or base plate size should be provided to resist this type
of failure.
The use of systems with frangible or sliding connections, or both, to the deck cantilevers
may be used to reduce the loads transferred to these cantilevers. There are some proprietary
systems that use these features.
C11.4 Expansion joints and end parapets
C11.4.1 Post and rail type barriers
Spigot and socket details should ensure continuity in lateral bending stiffness.
Joint details in metal railings should allow for removal and replacement of any damaged
section. Anchor bolts and other fixings cast into concrete should be designed to minimize
their damage when the attached post or railing has reached its plastic capacity.
Adequate expansion and contraction allowance in railing joints should be made for the
working movements. Thermal movements of metal rails will be much greater than those for
concrete parapets and superstructures.
Special end posts near expansion joints or barrier rail discontinuities should have
appropriate design strengths in the longitudinal direction to anchor the barrier. It is
suggested that this strength should be no less than the design strength of an intermediate
post in the transverse direction. This may be achieved by the combination of two
intermediate posts spaced close together.
C11.4.2 Rigid parapets
Where a significant width is required in parapet joints at expansion joints or other
discontinuities, consideration should be given to the use of recessed sliding cover plates to
ensure a smooth contact surface for traffic and to transfer shear across the gap.
C11.5 Pedestrian barriers
Pedestrian barriers should be detailed to provide safety and guidance to pedestrians. Design
details should avoid the use of sharp protrusions and offer smooth contact surfaces.

www.standards.org.au  Standards Australia


AS 5100.2 Supp 1—2007 40

C12 DYNAMIC BEHAVIOUR


C12.1 General
The dynamic interaction of bridges and moving vehicles is extremely complex, as is human
perception of and response to bridge movements. Certain levels of vibration may cause
physical discomfort particularly to pedestrians and occupants of stationary vehicles.
Vibration may also cause fatigue problems in some structures; however, this is covered by
the fatigue criteria specified in Clauses 6.9 and 8.7 of the Standard. Serviceability limit
state criteria to control unacceptable vibrations are not well developed and require further
research.
The difficulties of obtaining precise information on the dynamic loads and of finding
relevant acceptability criteria, means that the designer is unlikely to obtain solutions with
the confidence levels that apply to static problems.
Although AS 2670.1, Evaluation of human exposure to whole-body vibration—General
requirements, provides data on the evaluation of human exposure to whole body vibrations,
it also indicates that there are limits to its usefulness for the evaluation of disturbance due
Accessed by Brandon & Associates refreshed 170613 on 23 Oct 2019 [OBSOLESCENT] (Document currency not guaranteed when printed)

to vibration.
Vibration can usually be controlled by limiting the frequency of the fundamental mode of
vibration of the structure to a value markedly different from the frequency of the source of
vibration.
Alternatively, the structure may be suitably damped to reduce the magnitude of the
vibrations to an acceptable level.
Among the dynamic loads that may require consideration for the serviceability limit state
are—
(a) road or railway traffic on bridges with pedestrian walkways for the public;
(b) pedestrian traffic on road bridges and pedestrian bridges (Ref. 38);
(c) wind loads on structure (Ref. 39); and
(d) blasting (assuming a nearby quarry) (ACI 215-74) (Ref. 40).
C12.2 Road bridges
The vibration provisions for road bridges in the Standard are based on a similar approach to
the Canadian Highway Bridge Design Code 2000 (Ref. 7), modified for SM1600 loading
and verified by the findings of limited testing of Australian bridges.
The deflection of a bridge under the passage of a single heavy vehicle includes the
following components:
(a) The static or crawl speed deflection.
(b) Dynamic oscillations occurring at the natural frequency of the bridge.
Data from Canadian and Australian field investigations, involving vehicles with mass from
15 to 60 t, indicated an average dynamic increment of 12 to 15% of the static deflection.
Human response is primarily related to the vertical motion of the bridge. Research indicates
that acceleration is a reasonable measure of human response for bridges with a fundamental
natural frequency in the range 1 to 10 Hz, i.e., span lengths of about 10 to 100 m.
Vertical accelerations of approximately 1, 2 and 5 m/s2 [BS 5400—1978 (Ref. 41)]
respectively, are considered by most people to be very good, acceptable and unpleasant.
Reference may also be made to specialist literature and to AS 2670.1, which provides some
guidance on human response to whole-body vibration.

 Standards Australia www.standards.org.au


41 AS 5100.2 Supp 1—2007

Situations where traffic is congested, more than one lane is occupied or vehicles comprise
multi-axle, heavily loaded trucks, do not generally produce significant vibration effects in
bridges. This is because the probability of many axle suspension systems vibrating in phase
is remote. The situation most likely to produce significant vibration effects is where a
single vehicle is on the bridge. The Canadian Highway Bridge Design Code (Ref. 7) uses
the acceleration caused by a representative vehicle traversing the superstructure as the
serviceability limit state load for design of vibration for highway bridges. To simplify the
design process, accelerations were converted to equivalent static deflections, based on field
observations of the dynamic responses of bridges to traffic in a travelled lane. It was also
assumed that bridges with significant pedestrian use would be crossed by vehicles at low
speed (60 km/h or less).
C12.2.1 With walkways
The serviceability design vehicle for vibration is a single 0.7 (M1600 moving traffic load
without UDL).
The deflection limits for road bridges with walkways shown in Figure 12.2.1 of the
Accessed by Brandon & Associates refreshed 170613 on 23 Oct 2019 [OBSOLESCENT] (Document currency not guaranteed when printed)

Standard are based on the Canadian Highway Bridge Code (Ref. 7) provisions for bridges
with walkways and few pedestrians. Using the methodology of Wright and Walker (1971)
(Ref. 42) indicates that this is approximately equivalent to limiting accelerations to less
than about 3 m/s2 for most bridges. The allowable deflections are greater than in the
ABDC-1992 (HB 77.2—1996) (Ref. 1) because of the increased mass of the serviceability
design vehicle for vibration (previously a T44 truck) and the generally satisfactory
performance of existing Australian bridges.
Compliance with the deflection limits should generally ensure that structures perform
satisfactorily in respect to vibration effects.
Kerbs and parapets can have a significant effect on the deflection at a walkway, and may be
included in the structural model used to calculate deflections.
C12.2.2 Without walkways
Where a road bridge without a walkway complies with the deflection limits specified in
Clause 6.11 of the Standard, its vibration behaviour should generally be acceptable to
occupants of stationary or slow moving vehicles on the bridge. Where these limits are
exceeded, or where the structure is particularly important or unusual, a detailed assessment
of vibration effects should be carried out and specialist literature consulted.
C12.2.3 Detailed dynamic analysis
Refer to specialist literature.
C12.3 Railway bridges
Where railway bridges have walkways or platforms attached, which are intended for public
use, their dynamic behaviour should be controlled so as to prevent uncomfortable
vibrations.
The vibration characteristics of railway bridges not subject to pedestrian use may
nevertheless need to be investigated. Designers should note that acceptance criteria may
relate to trackwork or overhead electrification and its relationship to train performance.

www.standards.org.au  Standards Australia


AS 5100.2 Supp 1—2007 42

C12.4 Pedestrian bridges


When the repeated footfall of a pedestrian crossing a pedestrian bridge is near one of the
resonant frequencies of the structure, excitation, which approximates a steady state
vibration, can occur. In some instances, it can be alarming or objectionable to the user,
resulting in complaints to the authority. Therefore, pedestrian bridges should be checked
during design for undesirable vibration tendencies [Wheeler (1981) (Ref. 43)]. It can be
shown that a single 700 N pedestrian, walking at gaits in the range 1.75 to 2.5 footfalls per
second, produces a response that gives a good indication of such tendencies. Methods
available, which can readily model the pedestrian and calculate the response for comparison
with the acceptance criterion [(Wheeler) (1980) (Ref. 44)] are shown in Figure 12.4 of the
Standard.
Pedestrian bridges with resonant frequencies falling within the range of 1.5 to 3.5 Hz have
been found to have the greatest susceptibility to significant levels of vibration. Where
calculated frequencies are within this range, an investigation is recommended. For
pedestrian bridges with calculated resonant frequencies of 5 Hz and higher, it is highly
unlikely that vibration will be a problem.
Accessed by Brandon & Associates refreshed 170613 on 23 Oct 2019 [OBSOLESCENT] (Document currency not guaranteed when printed)

Where vibration is identified as a problem, a tuned-mass damper may provide an economic


solution. The assessment for any damping requirements should be made toward the
completion of construction, after the installation of railings and all restraints.
When investigating the resonant frequency of a pedestrian bridge using the method
described by Wheeler (1980) (Ref. 44), the response reduction factor should be applied with
caution to the maximum calculated displacement amplitude, especially when the critical
pedestrian frequency at which the amplitude response peaks is within the range of 2 to
3 footfalls per second.
Experiences with the Millennium Bridge in the UK have highlighted the importance of
designing to avoid or minimize synchronous lateral excitation.
Pedestrian bridges should be designed so as to avoid synchronous lateral excitation under a
specified crowd. Dallard et al (2001) (Ref. 45) and Dallard (2001) (Ref. 46) may be used as
a source of data on the magnitude of the excitation and limiting frequencies. This paper
notes the following:
(a) Synchronous lateral excitation need not be considered for modes with lateral
frequencies above 1.3 Hz.
(b) When calculating lateral frequencies, the proportion of the crowd mass considered to
act with the bridge mass is assumed to lie within the range of 0 to 30%. The value
adopted should be that causing the more onerous outcome.
(c) The value of k should be taken as 300 Ns/m over the full range of 0 to 1.3 Hz.
C12.5 Special structures
Refer to specialist literature.

C13 EARTH PRESSURE


C13.1 General
(No Commentary)

 Standards Australia www.standards.org.au


43 AS 5100.2 Supp 1—2007

C13.2 Surcharge loads from road traffic loads


The surcharge load specified in AS 5100.2 provides for the load effect from live load and
should be used in the design of all abutment walls and wing walls located within the
specified distance from the traffic. For high walls, the contribution of live load to the total
effects on the wall diminishes. This effect is simulated by the linear variation of surcharge
load from an equivalent fill height of 1.0 m to 0 m between the depths of 3 m and 8 m
below the top of the wall, as shown in Figure C13.2. Teng (1962) (Ref. 47) and Terzaghi
(1954) (Ref. 48) provide additional information.
There have been cases of damage to bridge abutments, designed with an approach slab,
arising from extremely high static and dynamic loads from construction traffic, prior to
placement of the approach slab. In view of the difficulties in controlling construction
traffic, it is considered that the live load surcharge should be included in the design of all
vehicular bridge abutments.
Accessed by Brandon & Associates refreshed 170613 on 23 Oct 2019 [OBSOLESCENT] (Document currency not guaranteed when printed)

FIGURE C13.2 RESULTING EARTH PRESSURE LOAD


DUE TO LIVE LOAD SURCHARGE

C13.3 Surcharge loads from railway loads


The Clause sets out requirements for the design of structures located under earth fill and for
the design of all abutment walls, wing walls, retaining walls and the like, located within the
specified distance from the rail load.
The method of determining rail surcharge loads is based on the simple concept of
permitting the surcharge load at the underside of sleepers to be distributed as appropriate
with increasing distance from the surcharge load to the structure.

www.standards.org.au  Standards Australia


AS 5100.2 Supp 1—2007 44

Additional information on methods of determining earth pressure on structures resulting


from rail traffic surcharge load are given in AREA Manual (1989) (Ref. 49).

C14 EARTHQUAKE FORCES


C14.1 General
The Clause is based on AS 1170.4—1993, Minimum design loads on structures—
Earthquake loads. While the general concepts in AS 1170.4—1993, such as the earthquake
risk, the influence of site (foundation) conditions and the structural system on earthquake
effects and the calculation of horizontal and vertical earthquake forces, are valid for all
structures types, many of the details in AS 1170.4—1993 relate to building structures only.
For this reason, earthquake effects to be considered in the design of bridge structures need
to be dealt with separately; however, where common sections of AS 1170.4—1993 can be
used directly, they are not repeated in the Standard and the design engineer is referred to
the relevant part of AS 1170.4—1993.
AS 1170.4—1993 is significantly different to its predecessor AS 2121—1979, The design of
Accessed by Brandon & Associates refreshed 170613 on 23 Oct 2019 [OBSOLESCENT] (Document currency not guaranteed when printed)

earthquake-resistant buildings, in its assessment of earthquake risk in Australia and its


philosophy for earthquake design. The old ‘zone’ map in AS 2121, which designated many
areas of Australia as seismically inactive, has been replaced by a contour map that defines
areas of uniform seismic risk. In the design approach, the ductility of the structure can now
be utilized, allowing a corresponding reduction in the earthquake design load.
In assessing the effects of earthquakes on bridge structures, either static or dynamic
analysis is used to calculate earthquake forces, depending on the bridge earthquake design
category (BEDC) and the dynamic structural response of the bridge.
C14.2 Limit state
Even in the most seismically active areas of the world, the occurrence of a design
earthquake is a rare event. The risk of structural and non-structural damage and plastic
deformations of bridge structures has to be offset against the increased cost associated with
design to resist earthquake forces in the elastic range.
The philosophy for earthquake design adopted in both AS 1170.4—1993 and the Standard is
to allow damage but to prevent collapse in the rare event of a very strong earthquake. The
bridge should be designed to encourage the formation of plastic hinges that will create a
collapse mechanism to dissipate the earthquake-induced forces. The required elastic
strengths for such an event may therefore be substantially reduced for design on the
provision of adequate ductility. In the design approach, the maximum elastic earthquake
forces are modified by the structural response factor to give equivalent ultimate limit state
earthquake design effects.
In Australia, given the levels of earthquake expected, a structure designed by this principle
will also usually satisfy the serviceability limit state, where for less severe earthquake, it is
required to remain in the elastic range, without experiencing significant damage.
C14.3 Bridge earthquake design category (BEDC)
Bridge structures are classified in one of four design categories in accordance with their
classification type, their location (level of earthquake risk) and site (foundation) conditions.
This category determines the earthquake design requirements.
The bridge classification recognizes the role many bridges would be required to play in a
post-earthquake situation as essential communication routes. In addition, for many bridge
structures, the consequences of earthquake damage may be of significance to other
structures or the public, even though the bridge itself may not be an important link.

 Standards Australia www.standards.org.au


45 AS 5100.2 Supp 1—2007

The acceleration coefficient (a) is a measure of the relative earthquake of a particular


location. It is an effective peak ground acceleration corresponding to return interval of
approximately 500 years and is expressed as a fraction of the gravity constant (g). While
this return interval is less than that defined for other bridge ultimate limit state design
loadings (typically 2000 years), it is considered appropriate to use the same basis for
determining the level of earthquake risk for bridges as for general structures in Australia.
The table and figures in AS 1170.4—1993 may therefore be used directly.
The site factor (S) is a measure of the effect of the foundation condition on the earthquake
motion and should be established from substantiated geotechnical data. The site factor
varies from 0.67 for rock to 2.0 for soft clays and silts. In determining the site factor, only
the material below the actual founding level needs to be considered. This is particularly
relevant for piled foundations, which may pass through soft soils, yet are founded in stiff
soil layers or rock. Soft soils tend to substantially modify earthquake ground motions over
those that the same earthquake would produce in solid rock. As a general rule, stiff
structures will tend to resonate on stiff ground whilst soft structures will tend to resonate on
soft soils. The Table in AS 1170.4—1993 may be used directly.
Accessed by Brandon & Associates refreshed 170613 on 23 Oct 2019 [OBSOLESCENT] (Document currency not guaranteed when printed)

C14.4 Requirements for earthquake design


The analysis and structural detailing requirements for bridges depend on the bridge
earthquake design category, the maximum span of the bridge and its dynamic structural
response.
The majority of Australian bridges will require, at most, a static analysis. Many short span,
low risk bridges will not need any special assessment for earthquake effects.
For special or important bridges, the road authority may elect to adopt the design
requirements for more severe BEDC where warranted.
Where the effects of earthquake forces need to be assessed, horizontal forces should be
considered in each major direction of the bridge structure. Vertical earthquake forces
become significant when the proportion of vertical loads due to live load effects relative to
those due to permanent loads becomes small. This is generally the case for longer span
bridges.
For bridge structures where more than one mode of free vibration contributes to the
dynamic response under earthquake loading, the equivalent static force approach may not
yield a good approximation of earthquake effects, and a dynamic analysis should be carried
out.
Also, when a bridge structure has a complicated or irregular distribution of mass or
stiffness, or both, a dynamic analysis should be carried out to more accurately assess the
distribution of the earthquake forces in the structures. Such cases include bridges where the
change in mass or stiffness, or both, along the length from one support to an adjacent
support exceeds 25% and curved bridges where the subtended angle, from abutment to the
other, at the centre of curvature is greater than 60°.
Bridges in the most severe earthquake design category need to be assessed by a dynamic
analysis method. In addition, it should be ensured that such bridges exhibit highly ductile
behaviour under the extreme collapse loadings typical of severe earthquakes.

www.standards.org.au  Standards Australia


AS 5100.2 Supp 1—2007 46

C14.5 Static analysis


The equivalent static force approach is a simplification that can be satisfactorily applied to
many structures. Essentially, it assumes that the horizontal and vertical acceleration
spectrums can be replaced by the quantity I(CS)/R f . Implicit in this approach is the
assumption that the first mode shape of the structure in a particular direction is a dominant
one and is uncomplicated. For example, a straight continuous deck on piers with bearings
would have a simple mode shape in a direction parallel to the deck (longitudinal flexure of
piers) and also a simple mode shape in a direction normal to the deck (transverse flexure of
piers and possible transverse flexure of the deck). Therefore, in instances where bridges
have high slenderness piers, significant curvature in plan (i.e., torsion is significant) or
significant variation in pier stiffnesses, it is recommended that the shape of several modes
be considered to determine whether the equivalent static force approach is appropriate. The
applicability of the method to bridges where the period of motion is long (say 3 s or more)
is also questionable. The equivalent static force approach should also be applied with some
caution to long bridges, since the ground motion towards each end may, in fact, be in
opposite directions.
Accessed by Brandon & Associates refreshed 170613 on 23 Oct 2019 [OBSOLESCENT] (Document currency not guaranteed when printed)

Earthquake forces should be assumed to be applied in any direction, but only as


independent loadings, for example, horizontal and vertical earthquake forces do not need to
be considered concurrently.
Because of the rarity of occurrence of a design earthquake, only permanent or long term
effects need to be considered in the calculation of earthquake forces.
A vertical earthquake acceleration component, equal to at least 50% of the maximum
horizontal force, needs to be included in the case of longer span bridges, where vertical
loads due to live load tend to decrease in significance. Both Bull (1972) (Ref. 50) and
Newmark and Rosenbleuth (1971) (Ref. 51) stress the importance of designing for this,
particularly where it is in the opposite sense to the gravity force.
Vertical earthquake forces are calculated using the same approach as for horizontal forces,
but using the period of the first dominant mode of free vibration (T) in the vertical direction
(typically corresponding to flexure of the deck).
The importance factor (I) recognizes the role bridges would be called on to perform in a
post-earthquake situation. Where they are a part of an essential communication route, they
should be able to function after natural disasters such as earthquakes. However, where
alternative routes may be available, the authority may elect to reduce the value of T to a
value between 1.25 and 1.0.
The earthquake design coefficient (C) is the factor that converts the ground acceleration
into a total equivalent earthquake force in a particular direction. It is a function of the
structure period of the dominant mode of free vibration in the direction under consideration.
The structure period (T) should be determined by a computer analysis or other recognized
theoretical approach. For bridges in BEDC-1 only, the approximate methods given may be
used if these are simpler.
The structural response factor (R f ) is a measure of the ductility and the capacity of the
structural system to resist earthquake-induced forces. It is essential the ratio of the forces
that would develop if the structure behaved entirely in the linear elastic range to the forces
that the structure could sustain with significant yielding under the design earthquake
effects. R f varies in accordance with the type of structural system that is available to resist
earthquake-induced forces. Effectively, R f , which is applied to the calculated elastic
earthquake forces, reduces the earthquake forces for which the structure should be designed
in proportion to its inherent ductility. Less ductile structural forms thus need to be designed
for greater earthquake forces.

 Standards Australia www.standards.org.au


47 AS 5100.2 Supp 1—2007

C14.6 Dynamic analysis


A dynamic analysis is carried out to provide a more accurate distribution of the earthquake-
induced forces and a better estimate of the response of the bridge structure’s natural modes
of vibration to earthquake loadings.
A dynamic analysis may take the form of a response spectrum analysis, using the modal
analysis technique or a time history analysis performed generally in accordance with
AS 1170.4—1993.
A time history analysis requires an input consisting of acceleration versus time record of an
earthquake. This method is the best simulation of the response of a structure to earthquake
effects, but generally requires more computing effort and a special procedure to select a set
of representative earthquake records appropriate to each specific site, which may not always
be available. Therefore, the response spectrum method is generally recommended for
dynamic analysis of conventional bridges.
The normalized response spectrum provided in AS 1170.4—1993 has been derived from an
ensemble of earthquake records for a number of different soil conditions. Most standard
Accessed by Brandon & Associates refreshed 170613 on 23 Oct 2019 [OBSOLESCENT] (Document currency not guaranteed when printed)

commercially available dynamic analysis software packages are able to utilize a response
spectrum in a modal analysis. The final structural response in each direction is the
aggregate of several modes of vibration, each considered separately under a corresponding
acceleration from the spectrum and combined to produce the total response. A sufficient
number of modes should be included in the response spectrum analysis so that at least 90%
of the structure’s mass (considered loads only) has been accounted for, in each direction
considered. The measure of this is often referred to as the mass participation factor.
C14.7 Structural detailing requirements for earthquake effects
C14.7.1 General
While not all bridges require special detailing for earthquake effects, it is good practice to
incorporate some detailing specific to earthquake effects, such as ensuring some degrees of
ductile behaviour and providing minimum horizontal load restraints, to minimize the effects
of unexpected earthquakes.
The subject of the earthquake design of bridges is dealt with very comprehensively in the
New Zealand National Society for Earthquake Engineering Bulletin (1981) (Ref. 52).
C14.7.2 Restraining devices
Care should be taken in detail design to ensure the integrity of the structure during an
earthquake. Whenever possible, the superstructure should be designed as continuous,
otherwise positive horizontal linkages should be provided between adjacent sections of the
superstructure at supports and hinges. Such provisions should cater for earthquake forces in
both the transverse and longitudinal directions. In meeting this requirement, it should be
remembered that restraining devices provide very little energy dissipation and will therefore
not eliminate the need for adequate ductile behaviour in supporting members.
The forces in restraining devices are very difficult to predict analytically because of the
non-linear behaviour created by expansion gaps. The Japanese Society of Civil Engineers
(1973) (Ref. 53) sets out methods by which the proportioning of earthquake forces between
the superstructure and the substructure may be determined. The horizontal force to be
resisted by restraining devices may be reduced, where applicable, by column or bearing
shears due to earthquake forces; however, it should not be less than the minimum lateral
restraint force specified in Clause 9 of the Standard.

www.standards.org.au  Standards Australia


AS 5100.2 Supp 1—2007 48

As earthquake forces are generally cyclic in nature and may result in significant reversals
and fluctuations, the component of friction often employed in the design of some anchorage
systems, such as bearings, cannot be relied upon under such loadings. Therefore, the
coefficient of friction between any material types should be taken as zero under earthquake
forces.
C14.7.3 Provision of horizontal movements
Bearings (including sliding and elastomeric types) and deck joints should be able to
accommodate the horizontal movements induced by earthquake forces.
Where sliding bearings are desirable, the stiffness of the support in resisting horizontal
earthquake forces may be ensured by the use of stops that are outside the normal range of
movement. Such devices may also be used to limit excessive movements, which may be
generated under earthquake forces only, so that conventional bearings and deck joints can
still be used.
C14.7.4 Soil behaviour
Accessed by Brandon & Associates refreshed 170613 on 23 Oct 2019 [OBSOLESCENT] (Document currency not guaranteed when printed)

The behaviour of soil under earthquake forces should not be overlooked, particularly in
large earth-retaining structures. A good deal of information and design methods are
available [see Berrill and Elms (1978) (Ref. 54), Japanese Society of Civil Engineers (1973)
(Ref. 53), and Ferriot and Forrest (1977) (Ref. 55)]. In particular, the Mononobe-Okabe
method may be used for determining equivalent static fluid pressures for earthquake forces
on retaining walls, and is described in detail by Seed and Whitman (1970) (Ref. 56), and
Richards and Elms (1979) (Ref. 57).
C14.7.5 Ductile behaviour
C14.7.5.1 General requirements
Structures in which members and their connections can sustain a number of load reversing,
post-yield deflection cycles whilst maintaining a substantial proportion of their initial load
carrying capacity are defined as ductile structures. Ductility ratio is often used as a measure
of this attribute. The ductility of a member or structure is typically defined as the ratio of
the deformation at failure to the deformation at yield, and is a measure of its ability to
undergo large deformations prior to collapse without a significant reduction in the structural
strength. Ductility is an essential design requirement for a structure to behave satisfactorily
under severe earthquake excitation. The ductility demand of a structure under such loadings
is dependent on the construction material, the design elastic strength and the structural
system.
Regarding design for ductility, Newmark and Rosenbleuth (1971) (Ref. 51) state that by
providing high ductility, we incidentally produce a structure capable of large scale
redistribution of stresses; and this in itself lowers the probability of failure under
unforeseen conditions of loading or oversights in design. Ordinarily, adequate ductility can
be incorporated at quite low cost. In steel, timber and pre-cast concrete, ductility requires
relatively strong connections and weak members. In reinforced concrete, it calls for a
higher capacity in diagonal tension than in bending. In all types of structures, it calls for a
smaller probability of inelastic buckling than of failure in tension, and it requires that very
special attention be given to detail.
In bridges where the piers, for example, are designed so as to be capable of dissipating
energy by the formation of ductile plastic hinges, sufficient strength capacity should be
provided elsewhere in the structure to ensure that the chosen energy dissipating mechanism
does develop in the event of a severe earthquake. The chosen system should involve plastic
hinges in the piers rather than the foundations because of the greater accessibility for
inspection and repair.

 Standards Australia www.standards.org.au


49 AS 5100.2 Supp 1—2007

The importance of designing for ductile behaviour and avoidance of brittle failures in
concrete, in particular, should be recognized. Park and Blakeley (1978) (Ref. 58)
recommended that there be a suitable hierarchy of failure of the structural and foundation
members that will preclude a brittle collapse. This may be achieved by appropriate margins
of strength between non-ductile and ductile members, and with attention to detail. The New
Zealand Highway Bridge Design Brief (1973) (Ref. 59) recommends the provision of
additional shear capacity to preclude brittle failure in that mode. In plastic hinge regions
subjected to reversed flexure, little reliance can be placed on shear transferred by aggregate
interlock or via the compression zone of the concrete due to cyclic degradation of these
mechanisms. Therefore, it is recommended that, in these regions, shear reinforcement be
provided to carry the total shear force by dowel and truss action.
The provisions for lateral reinforcement at potential plastic hinge locations are to provide
confinement of the concrete and prevent buckling of the longitudinal reinforcement.
It is also recommended that splices in longitudinal reinforcement be located well away from
potential plastic hinge areas to reduce the likelihood of a brittle failure in the splice region.
Accessed by Brandon & Associates refreshed 170613 on 23 Oct 2019 [OBSOLESCENT] (Document currency not guaranteed when printed)

For bridges in BEDC-4, a special assessment of the post-elastic behaviour of the structure is
necessary to ensure that the preferred collapse mechanism is defined. Minimum ductility
requirements are given to ensure that potential plastic hinges can develop and sustain the
required loads. These requirements are based on New Zealand earthquake design
requirements.
The limit on the design axial compression force in a member at plastic hinges is necessary
as axial compression can significantly reduce the curvature ductility of a section. Spalling
of concrete cover in members with axial compression occurs at an earlier stage than in those
without, although the provision of transverse confining reinforcement markedly improves
ductility. A reinforced concrete section with a large difference between the tensile and
compression reinforcement area will also exhibit reduced ductility. In prestressed concrete
members, it is necessary to provide a minimum area of non-prestressed reinforcement as
typically this steel behaves in a more ductile manner than prestressing steel, having a
clearly defined yield limit and a significant capacity to undergo non-linear deformation.
Where a post-elastic analysis does not consider geometric non-linear effects, it is
recommended that the maximum horizontal deflection of column members not exceed
H/150, where H is the total structure height. This will limit Pδ effects in column members.
C14.7.5.2 Pile to pile cap connection
(No Commentary)

C15 FORCES RESULTING FROM WATER FLOW


C15.1 General
At locations where the road and bridge system intersects a catchment area, bridges are
liable to be affected by flooding. Bridges across large areas of water, estuaries and open sea
on the other hand will be subject to tidal and wave action. For the latter effects, the design
engineer should consult the extensive literature on the design of offshore structures.
Hogben et al (1977) (Ref. 60) provides an excellent working guide for loads resulting from
waves and Leonard et al (1981) (Ref. 61) gives a review of the literature.
Designers are reminded that, in estimating water flow effects caused by the presence of a
bridge structure, if scour is possible then the following conditions should be investigated:
(a) No scour, when determining flood loads.
(b) Maximum scour, which will generally prove critical for structural analysis.
(c) Minimum or no scour, which is critical for backwater effects.

www.standards.org.au  Standards Australia


AS 5100.2 Supp 1—2007 50

C15.2 Limit states


C15.2.1 Ultimate limit states
In instances where the design flood for ultimate limit state will substantially overtop the
bridge superstructure, intermediate stages in the flood height should be investigated. This is
important because the maximum water forces on the superstructure can occur when it is just
overtopped. Investigation of intermediate stage heights may preclude the need for
investigation of the design flood for the 2000 year ultimate limit state return interval.
C15.2.2 Serviceability limit states
(No Commentary)
C15.3 Forces on piers due to water flow
C15.3.1 Drag forces on piers
In bridge structures subjected to water flow, the major loads are sustained by the piers.
There is a scarcity of information on the calculation of water flow forces on piers. The
Accessed by Brandon & Associates refreshed 170613 on 23 Oct 2019 [OBSOLESCENT] (Document currency not guaranteed when printed)

forces take the form of drag, parallel to the direction of flow, plus transverse lift, when the
flow is skewed to the plane of the pier, and sometimes fluctuating transverse forces when
the shape of the pier is bluff and produces vortices in the flow. Water flow forces tend to
become significant when piers are tall and slender with flows of considerable depth and
velocity.
C15.3.2 Lift forces on piers
Lift type forces have generally been ignored by codes in the past, but these can be large for
plate or wall type piers and for submerged decks that are super-elevated. Since there is no
guarantee that, under flood conditions, flow direction will not be angled to pier axes, design
engineers should employ pier shapes that are insensitive to angled flow. Consideration
should be given to flood forces being applied to piers and abutments during construction
and before the deck structure has been installed. The subject is discussed at some length by
Apelt (1965) (Ref. 62), and Apelt and Isaacs (1968) (Ref. 63). Little investigation seems to
have been devoted to water forces on different pier shapes; however, these loads will
generally be less critical than debris loading.
C15.4 Forces on superstructures due to water flow
Jempson and Apelt (1995), (1997), (1997) (Refs 64, 65 and 66) and Jempson (1995), (2000)
(Refs 67 and 68) have reported on the subject in considerable detail. Flood loadings on
bridge superstructures were measured in a comprehensive laboratory program for ranges of
flood conditions and geometric arrangements likely to be encountered in practice. The
forces and moments were measured on the six scale models of superstructures shown in
Figure C15.4.
The effect of the Froude number (F), degree of submergence (SR), and proximity of the
superstructure to the bed (Pr) on the forces and moments were investigated in a parametric
study. The forces and moments on superstructures were found to be strongly dependent on
these parameters and, so, for some flow and geometric conditions, the load coefficients
have been significantly changed from the ABDC-1992 (HB 77.2—1996) (Ref. 1). Also
studied were the effects of turbulence intensity, superelevation, and skew.
The coefficients presented in the Standard are applicable to prototype bridges of similar
geometry to the first four model types in Figure C15.4. Design information for protoypes of
similar geometry to the truss and box girder models is given in Jempson (1995) (Ref. 67).

 Standards Australia www.standards.org.au


51 AS 5100.2 Supp 1—2007

The ABDC-1992 (HB 77.2—1996) (Ref. 1) specified that a C d of 2.2 be adopted in absence
of more exact estimates. Jempson and Apelt (1995), (1997), (1997) (Refs 64, 65 and 66)
and Jempson (1995), (2000) (Refs 67 and 68) found that the Cd was strongly dependent on
S R and P r. These relationships are incorporated into the recommended values of Cd shown in
Figure 15.4.2(A) of the Standard. The strong dependence of C d on P r is due to wake
blockage. This phenomenon is described in detail in Refs 64 to 68.
Historically, the drag force has been applied at the mid-height of the superstructure to
calculate the overturning moment. Although the mid-height of the superstructure is unlikely
to be the correct line-of-action of the drag force, any error would not be a significant
proportion of the total length of the lever arm; however, historically the hydrodynamic lift
force has not been included in the calculation of the overturning moment. Jempson and
Apelt (1995), (1997), (1997) (Refs 64, 65 and 66) and Jempson (1995), (2000) (Refs 67
and 68) reported significant hydrodynamic lift forces that are eccentric to the centre-line of
the cross-section, i.e., additional overturning moment is generated that should be accounted
for in the design. The moment presented in Clause 15.4.4 of the Standard accounts for both
the vertical eccentricity of the drag force relative to the soffit as well as the lateral
Accessed by Brandon & Associates refreshed 170613 on 23 Oct 2019 [OBSOLESCENT] (Document currency not guaranteed when printed)

eccentricity of the lift force relative to the centre of the bridge.


The moment on a superstructure (M g ) (see Clause 15.4.4 of the Standard) accounts for the
eccentricity of the drag and lift forces. M g and F d are used to resolve an equivalent line of
action above the girder soffit for F d. This is not the true line of action for F d because it
incorporates the moment generated by the eccentricity of the lift force; however, applying
F d along this line will correctly calculate the overturning moment to be resisted by the
bridge. An alternative approach is to apply F d at the level of the girder soffit and add M g .
Upper bound and lower bound lines are shown in Figure 15.4.3 of the Standard for the
superstructure C L . The lower bound line was set at approximately the most negative C L
measured for the four model types, and the upper bound line was set at approximately the
least negative C L . C L was derived from hydrodynamic forces only, so the designer should
make allowance for buoyancy forces determined using the unrestricted tailwater level to
calculate the displaced volume: The unrestricted tailwater level is the downstream water
level that would occur without the bridge in place.
The lower bound line should be used for design cases where a downward force is
significant, e.g., compressive loading on piles or footings, girder design and deck design.
These design loads should be a combination of M g and the downward lift force distributed
uniformly across the bridge deck. The upper bound line should be used when considering
the stability of the bridge in overturning. The use of the upper bound line is generally
conservative. If overturning is critical in the design, it is recommended that the designer
review the data reported by Jempson (1995) (Ref. 67) to take greater advantage of the
stabilizing effect of the downward lift.
The designer should consider a range of floods up to the design event, not just the design
flood. For example, if the design flood has an S R of 3.5, the bridge will still be subject to
larger downward lift force at SR equals 1.5 as the flood rises.
The design velocity is that of the approach flow at a depth equivalent to the mid-height of
the superstructure for the design flood event under consideration. For most cases the surface
velocity should be adopted, but when the superstructure is deeply submerged, the average
stream velocity rather than the higher surface velocity may be more appropriate. The advice
of a suitably experienced hydraulics engineer should be obtained before adopting the
average stream velocity.

www.standards.org.au  Standards Australia


AS 5100.2 Supp 1—2007 52

Jempson (1995) (Ref. 67) undertook limited testing on the model box girder bridge with a
25 skew to the flow. Displacement of the model along the centre-line of the bridge when the
model was skewed to the flow indicated loading transverse to the stream, but the testing
apparatus was unable to measure this load. The drag coefficient was 30% lower than that
for the model square to the flow under the same condition. Further research is required, but
in the interim it may be appropriate to apply a load equivalent to 30% of the drag force
along the longitudinal centre-line of the bridge. The drag force should not be reduced. The
transverse load would be applied in the direction of the upstream abutment.
Accessed by Brandon & Associates refreshed 170613 on 23 Oct 2019 [OBSOLESCENT] (Document currency not guaranteed when printed)

DIMENSIONS IN MILLIMETRES

FIGURE C15.4 MODEL BRIDGE SUPERSTRUCTURES

 Standards Australia www.standards.org.au


53 AS 5100.2 Supp 1—2007

C15.4.1 General
(No Commentary)
C15.4.2 Drag force on superstructures
(No Commentary)
C15.4.3 Lift force on superstructures
(No Commentary)
C15.4.4 Moment on a superstructure
(No Commentary)
C15.4.5 Loads on superstructures with superelevation
(No Commentary)
C15.5 Forces due to debris
Accessed by Brandon & Associates refreshed 170613 on 23 Oct 2019 [OBSOLESCENT] (Document currency not guaranteed when printed)

Real debris will accumulate against a bridge superstructure or pier in an almost infinite
combination of geometry (size and shape), roughness, and porosity, from flood to flood and
river to river, and the geometry, roughness, and porosity will be dynamic. The factors that
contribute to the form of the debris mat include the catchment vegetation, land uses, and the
flow characteristics of the flood. A debris mat might form in the following manner:
(a) A large tree becomes wedged against the superstructure or pier, or both.
(b) Other debris interlocks with the branches of the tree and so the mat increases in size.
(c) Smaller debris such as small brushes, grass, leaves, paper, clothing and the like, lodge
in the mat thereby decreasing the porosity of the mat. It would be reasonable to
assume that no significant debris mat would form in a catchment with no trees. In
catchments where there is clearing of forests, it is more likely that debris would be
transported downstream than in a virgin catchment.
Field observations of debris mats are limited because of the difficulty in accessing bridges
during floods. In the United States superstructure debris mats were observed to approximate
a wedge shape in catchments where logging is taking place and the felled and stripped trees
have been carried down a river until being trapped by a bridge.
Jempson and Apelt (1997) (Ref. 66) and Jempson (2000) (Ref. 68) have reported on the
subject in considerable detail. Drag forces from debris loadings on bridge superstructures
and piers were measured in a comprehensive laboratory program for ranges of flood
conditions and geometric arrangements likely to be encountered in practice. The drag forces
generated by debris mats wedged against model superstructures were measured, as were the
drag forces generated by debris mats wedged against piers.
Two debris mat geometries were used in the laboratory to provide realistic upper and lower
limits. They were a flat debris model, which simulates a tree lodged against the pier or
superstructures, and more streamlined wedge- and cone-shaped models, which simulate the
debris observed in the United States.
The first 5 model superstructures shown in Figure C15.4 and three pier types, a twin
column, a blade, and pile bent were used for this investigation.
The effects of F, SR , P r and flow depth on the drag force were investigated in a parametric
study. The drag force on superstructures in combination with a debris mat were found be
strongly dependent on the F, P r and the geometry of the debris mat. The drag forces on
piers in combination with a debris mat were found to be strongly dependent on the flow
depth, velocity and the geometry of the debris mat.

www.standards.org.au  Standards Australia


AS 5100.2 Supp 1—2007 54

The recommended drag coefficients are an upper bound and are based on the flat debris
model. Designers are referred to Jempson (2000) (Ref. 68) for design information if a
wedge-shaped debris mat is considered more appropriate.
The design velocity is that of the approach flow at a depth equivalent to the mid-height of
the debris mat for the design flood event under consideration. For most cases the surface
velocity would be adopted.
The drag coefficients are for debris mats in combination with the superstructure or pier as
per the laboratory testing, hence the requirement of Clause 15.5.4 of the Standard that the
debris drag force be not used concurrently with water flow forces, except for a lift force to
be used in determining the structures resistance to overtopping.
C15.5.1 Depth of debris mat
(No Commentary)
C15.5.2 Debris acting on piers
(No Commentary)
Accessed by Brandon & Associates refreshed 170613 on 23 Oct 2019 [OBSOLESCENT] (Document currency not guaranteed when printed)

C15.5.3 Debris acting on superstructures


(No Commentary)
C15.5.4 Calculation of debris load
(No Commentary)
C15.6 Forces due to log impact
(No Commentary)
C15.7 Effects due to buoyancy and lift
(No Commentary)

C16 WIND LOADS


C16.1 General
A wind-sensitive structure is one in which additional loads occur as a result of the dynamic
interaction of the wind and the structure. Suspension bridges and long-span cable-stayed
bridges generally fall into this category.
C16.2 Design wind speed
C16.2.1 General
(No Commentary)
C16.2.2 Average return interval
The serviceability design wind speed corresponds to a 20 year return interval for wind in
combination with permanent effects.
It should be noted that, for wind-sensitive structures, the critical wind speeds will generally
be less than the 20 year return interval design wind speed. For wind in combination with
traffic load, a design wind speed of 35 m/s has been selected because, above this wind
speed, traffic movement is virtually impossible. This wind speed is applicable to all
localities in Australia. As congested traffic would be rare at wind speeds of 35 m/s, and as a
high ultimate design wind speed is to be used, the effect of wind on the traffic load need not
be considered.

 Standards Australia www.standards.org.au


55 AS 5100.2 Supp 1—2007

At ultimate limit states, the design wind speed corresponds to a return interval of
2000 years. AS/NZS 1170.2, Structural design actions—Wind actions, defines the ultimate
limit state design wind speed for a return interval of 1000 years for various locations. This
may be extrapolated for a 2000 year return interval as specified in AS/NZS 1170.2.
C16.3 Transverse wind load
C16.3.1 Calculation of transverse wind load
(No Commentary)
C16.3.2 Area of structure for calculation of transverse wind load (A t)
The area At includes the superstructure only, and the effect of wind on the traffic load need
not be considered.
C16.3.3 Drag coefficient (Cd)
Drag coefficients should be calculated for the following:
(a) Drag coefficient for all superstructures with solid elevation The depth of
Accessed by Brandon & Associates refreshed 170613 on 23 Oct 2019 [OBSOLESCENT] (Document currency not guaranteed when printed)

superstructure (d), including solid parapet if applicable, is shown in


Figure C16.3.3(A). The types of superstructure for which Figure C16.3.3(A) is
applicable are shown in Figure C16.3.3(B).
For structures such as shown in Figure 16.3.3(C), wind tunnel tests would be required
to determine accurate values of Cd .
For superstructures separated horizontally by an air gap of less than 1 m, and which
are supported by a common substructure, the substructure need only be designed
assuming a width of structure equal to the total width.
(b) Aerodynamic shape factor for truss girder superstructures The wind force may need
to be calculated for wind in a direction other than normal to the bridge centre-line.
(c) Drag coefficients for beams during erection Drag coefficients for circular cylinders
with high aspect ratio are given in AS/NZS 1170.2.

FIGURE C16.3.3(A) DEPTH OF SUPERSTRUCTURE (d) TO BE USED


FOR CALCULATION OF C d

www.standards.org.au  Standards Australia


AS 5100.2 Supp 1—2007 56

FIGURE C16.3.3(B) TYPICAL SUPERSTRUCTURES FOR WHICH FIGURE C16.3.3(A)


IS APPLICABLE
Accessed by Brandon & Associates refreshed 170613 on 23 Oct 2019 [OBSOLESCENT] (Document currency not guaranteed when printed)

FIGURE C16.3.3(C) TYPICAL SUPERSTRUCTURES THAT WOULD REQUIRE WIND


TUNNEL TESTS

C16.4 Longitudinal wind load


Generally, the longitudinal component of wind load on the superstructure is insignificant in
comparison with longitudinal traffic forces; however, for certain structures, such as tall,
balanced cantilever structures, this component should not be neglected during the
construction stages.
C16.5 Vertical wind load
The design vertical wind load for angles of inclination of the wind to the superstructure of
less than 5° has been calculated assuming a lift coefficient of 0.75.
C16.6 Wind on railway live load
Wind loads on railway live load need to be considered because of the realistic possibility of
trains being located on bridges during extreme wind events, and due to the fact that, in
many cases, a train will present a large surface area to the prevailing wind. This is in
contrast to highway loading where vehicles will rarely be present on a bridge during
extreme wind events, and because most highway vehicles will be cars, which would present
only a small area to the wind.

 Standards Australia www.standards.org.au


57 AS 5100.2 Supp 1—2007

Wind loads on a structure plus live load at the ultimate limit state are conservative. The
2000 year return interval wind, which is specified for the ultimate limit state, is applicable
to the bridge structure alone, but has been specified in the Standard for acting on live load
also to simplify its application. For the majority of railway structures this will not present a
problem, because the total wind load for short spans will generally be less than the
minimum lateral restraint capacity specified in Clause 9.
For large structures where the wind loads on live load are significant, a reduced wind speed
may be used for wind on live load at the ultimate limit state. The use of such a reduced
speed should be related to the specified operating conditions for the railway. Train
operations will generally cease at wind speeds above approximately 40 m/s because of
problems with flying debris.

C17 THERMAL EFFECTS


C17.1 General
(No Commentary)
Accessed by Brandon & Associates refreshed 170613 on 23 Oct 2019 [OBSOLESCENT] (Document currency not guaranteed when printed)

C17.2 Variation in average bridge temperature


The values for maximum and minimum shade air temperature given in Table 17.2(1) of the
Standard are based on meteorological data for extreme temperatures over a recording period
of approximately 15 years. For each of the 107 meteorological districts across Australia,
which have been defined by the Bureau of Meteorology, averages of the highest maximum
and lowest minimum temperatures recorded at stations within these districts were
determined. Analysis of these results indicated that simplification into the broad regions
given in Table 17.2(1) was warranted for design recommendations. The values for
maximum shade air temperature produce maximum average bridge temperatures similar to
those specified in the NAASRA Bridge Design Specification (1976) (Ref. 23). Those for
minimum shade air temperature result in minimum average bridge temperatures slightly
higher than previously given.
The recommendations for average bridge temperature given in Table 17.2(2) of the
Standard are based on the evaluation of measured temperature distribution in typical
concrete box girder bridges. The values do not appear to be as extreme as requirements
adopted by European codes since, in Australia, extreme weather conditions usually exist for
relatively short durations. For example, the Lake Ginninderra bridge in Canberra, with a
frozen deck, associated with surface temperatures of −3°C as shown in Figure C17.2(A),
has a minimum average temperature of 2.4°C. This is because the frozen condition occurs
in cycles with daytime thawing; however, measurements have shown that bridge
superstructures in Britain can have a relatively uniform low temperature.
Where sections are shallow more extreme maximum and minimum values may occur,
whereas for relatively deep sections the heat sink effect of concrete remote from the
surfaces could reduce the range of average bridge temperatures. Figure C17.2(B) shows the
extent of relatively unaffected web concrete, for a range of section depths, during a short
term temperature cycle.
For the case of minor structures, the range of displacements determined from the
appropriate shade and average bridge temperatures, rather than the temperatures
themselves, should be assessed and, where possible, increased to allow for limited
supervision and control of setting bearings and deck joints.
Average temperature effects for superstructures consisting of a concrete deck on steel
girders should also be considered during construction, prior to placement of the concrete
deck slab. Thermal movements of the steel girders alone may be a problem, particularly for
curved girders.

www.standards.org.au  Standards Australia


AS 5100.2 Supp 1—2007 58

DIMENSIONS IN MILLIMETRES
Accessed by Brandon & Associates refreshed 170613 on 23 Oct 2019 [OBSOLESCENT] (Document currency not guaranteed when printed)

FIGURE C17.2(A) EXAMPLE OF TEMPERATURE DISTRIBUTION

DIMENSIONS IN MILLIMETRES

FIGURE C17.2(B) TYPICAL BOX GIRDER TEMPERATURE VARIATIONS

 Standards Australia www.standards.org.au


59 AS 5100.2 Supp 1—2007

C17.3 Differential temperature


Differential temperature stresses may include primary and secondary (or continuity) effects.
These stresses need to be evaluated for both ‘hot top’ and ‘cold top’ conditions (positive
and negative temperature differentials), to determine their effect in addition to flexural and
axial stresses from other loadings. To enable a single, non-transient, thermal stress analysis
to be used, the design effective vertical temperature gradients shown in Figure 17.3 of the
Standard have been derived from actual critical temperature distributions in typical
sections.
In the case of concrete superstructure types subject to cyclic thermal stress reversals, the
effect of creep tends to produce a null stress state with a positive temperature gradient of
about 5°C. On the other hand, a zero gradient is produced in the case of a stationary thermal
stress state. Therefore, the consequence of the cyclic nature of the stress state is to
effectively reduce hot top temperature gradients, but increase those derived from cold top
conditions. This effect is discussed in detail by Thomas (1980) (Ref. 69). Figure C17.3
shows typical cyclic variations in temperature measured in a box girder cross-section.
Accessed by Brandon & Associates refreshed 170613 on 23 Oct 2019 [OBSOLESCENT] (Document currency not guaranteed when printed)

Creep effects on cyclic thermal stresses in steel composite sections are not as significant as
in full concrete sections. Therefore, the relevant design effective vertical temperature
gradients recommended for Type 3 superstructures in Figure 17.3 of the Standard show
increased hot top and reduced cold top requirements. Significant differentials through the
depth of steel components have not been measured in field studies.
The recommended design effective vertical temperature gradients shown in Figure 17.3 of
the Standard do not include an allowance for the effects of bituminous surfacing. For
thicknesses of surfacing up to about 50 mm there would be no significant effect on the
design gradient since the increase in absorption is approximately compensated by the
insulation effect of the surfacing; however, with greater thicknesses, the heat sink and
insulation effects could significantly reduce the temperature differential
(see BS 5400-2—1978) (Ref. 41).
Transverse differential temperature gradients across a superstructure may result from the
heating up of one side of the superstructure in relation to the other. In structures with
statically indeterminate lateral restraints, this effect should be included as a thermal load on
the restraint and lateral moments in the superstructure. A transverse linear temperature
differential of 15° should be allowed for, or else a thermal response analysis may be
needed.
The insulating effect of ballasted track will reduce the differential temperatures in railway
bridge decks.

www.standards.org.au  Standards Australia


AS 5100.2 Supp 1—2007 60
Accessed by Brandon & Associates refreshed 170613 on 23 Oct 2019 [OBSOLESCENT] (Document currency not guaranteed when printed)

FIGURE C17.3 CYCLIC TEMPERATURE VARIATIONS

C17.4 Limit states


The ultimate load factors have been taken from AS/NZS 1170.1.
Due to the strain compatibility nature of thermal stresses, for normal concrete
superstructures, cracking of sections at strength and stability, or serviceability loads causes
a reduction in the stiffness of these members and may lead to a redistribution and reduction
of thermal stresses.

 Standards Australia www.standards.org.au


61 AS 5100.2 Supp 1—2007

There is no general rule defining the manner in which thermal effects should be treated for
all material types at each limit state. Rather, each structure should be assessed
independently, considering both the material characteristics and structural behaviour.
Priestly and Buckle (1979) (Ref. 70) give comprehensive details concerning the effects of
thermal loadings on concrete sections; however, much research is still required. In the
absence of a more thorough investigation, the following guidelines may be helpful:
(a) For variation in average bridge temperature At serviceability and ultimate loadings,
all effects that are applicable to the structural form may be considered, unless
specifically excluded in AS 5100.5 or AS 5100.6.
(b) For differential temperatures—
(i) Reinforced and partially prestressed concrete At serviceability loadings, the
section is usually cracked; however, both primary and secondary effects should
be calculated using the uncracked section properties. The secondary
(continuity) effects may subsequently be reduced by the factor k, where k is the
ratio of neutral axis depth at the critical section under the serviceability loading
Accessed by Brandon & Associates refreshed 170613 on 23 Oct 2019 [OBSOLESCENT] (Document currency not guaranteed when printed)

(excluding thermal effects) to the overall section depth, and is greater than or
equal to 0.5.
At ultimate loadings, primary stresses may be neglected and secondary effects
calculated as above, with k equals 0.5. Web tensile stresses may also influence
the ultimate shear capacity and should be considered.
(ii) Fully prestressed concrete At serviceability loadings, the section is, by
definition, uncracked and both, primary and secondary stresses should be
included fully and calculated using uncracked section properties.
At ultimate loadings, the behaviour is similar to that of partially prestressed
sections.
(iii) Steel and steel composite At serviceability loadings (for all members) and
ultimate loadings (for slender members), both primary and secondary stresses
should be included (see AS 5100.6).

C18 SHRINKAGE, CREEP AND PRESTRESS EFFECTS


C18.1 Shrinkage and creep effects
In unrestrained concrete components and determinate concrete structures, the primary effect
of shrinkage and creep is dimensional change and deflection. The effects of such
deflections on the serviceability of the structure, including bearings and expansion joints,
should be assessed.
In restrained concrete components and indeterminate concrete structures, the dimensional
changes can cause significant stresses and redistribution of loads and reactions. This is
particularly important in composite structures with steel or precast concrete beams and a
cast-in-place concrete deck slab.
Tensile strains can exceed the tensile strength of concrete. Sufficient reinforcement should
be provided and properly distributed to control cracking.
In structures sensitive to creep and shrinkage effects, a more accurate estimate of the creep
and shrinkage properties of the concrete to be used should be obtained. Alternatively, the
possible range of properties should be determined and allowed for in the design.
Creep and shrinkage effects should be treated as permanent effects. The structure should be
checked for both early age (no creep or shrinkage) and long-term conditions (full creep and
shrinkage).

www.standards.org.au  Standards Australia


AS 5100.2 Supp 1—2007 62

C18.2 Prestress effects


The prestressing of restrained components and indeterminate structures usually produces
parasitic (secondary) moments, shears and reactions, which can form a significant load on a
structure. These effects should be included in both ultimate and serviceability limit state
loads.
The magnitude and distribution of parasitic effects are dependent upon the relative stiffness
of the components of the structure, e.g., beams of varying cross-section and structures
where piers are integral with the superstructure. These effects are also affected by cracking
at ultimate limit states, because when the response of a structure becomes non-linear the
principle of superposition of load effects is no longer valid and, in ductile structures,
redistribution of moments can occur. It is difficult to estimate the magnitude of these
parasitic effects, as the total effects of all loadings should be determined by a refined
incremental analysis; however, it is usually adequate to adopt the values of parasitic effects
calculated assuming uncracked sections and using an ultimate load factor of 1.0.
Similar comments apply to partially prestressed structures that may crack at serviceability
Accessed by Brandon & Associates refreshed 170613 on 23 Oct 2019 [OBSOLESCENT] (Document currency not guaranteed when printed)

limit states.
Both primary and secondary prestress effects should be included as permanent effects, and
the structure should be checked considering prestress with zero time-dependent losses and
prestress with full final losses.
Consideration of dead loads and prestress at transfer can be a critical design condition,
particularly for shear, and should be checked as a strength limit state. Maximum prestress
should be considered, with zero time-dependent losses, ungrouted tendons and an ultimate
load factor of 1.0 for all prestress effects. The ultimate load factors for unfavourable and
favourable dead loads are consistent with those used in other loading combinations.

C19 DIFFERENTIAL MOVEMENT OF SUPPORTS


C19.1 Differential settlement effects
The geotechnical investigation should include estimates of settlement over the various parts
of the bridge foundation. Significant differential settlements can occur whenever
foundations are not taken to rock.
Where embankment fills are placed over compressible substrata, abutment piles can suffer
substantial horizontal forces and deflections as well as downward forces during
consolidation, i.e., negative friction.
Differential settlement effects should be included in ultimate design loads where any part of
the structure has negligible plastic capacity, for example an over-reinforced concrete
section or a slender steel section, because of the inability of such sections to redistribute
action effects at the strength limit state.
The most common differential settlement problem with railway bridges occurs when simply
supported spans of less than 20 m long are strengthened for increased loading by adding
intermediate supports (converting them to two span continuous), where foundations are not
taken to rock. A full cycle of tension is applied to the anchor bolts by each vehicle from
increased uplift under differential settlement. Under this action, the anchor bolt nuts can
loosen, or fracture from fatigue, causing spans to rock on the intermediate supports. Once
this occurs, high dynamic load will be applied to the end reactions as each vehicle crosses
the span. This additional dynamic load can be sufficient to cause local failure of the
foundation material and rapid differential settlement, resulting in further increases in
dynamic load.

 Standards Australia www.standards.org.au


63 AS 5100.2 Supp 1—2007

C19.2 Mining subsidence effects


When an underground seam of coal is mined over a wide area, the immediate roof collapses
and material flows into the void. Movement is transferred through successive strata to the
surface resulting in the formation of a subsidence trough or bowl. The ground may not
necessarily crack.
The main features of subsidence are vertical displacements, changes in ground slope and the
development of surface strains, both tension and compression. The values vary according to
the extraction geometry, the geology and mining practice.
Estimates of the likely movements should be made after obtaining mining information from
the appropriate statutory authority. The National Coal Board (1975) (Ref. 71), Frankham
and Mould (1980) (Ref. 72), and Sims and Bridle (1966) (Ref. 73) give guidance on the
calculation of ground movements.
In designing and detailing a suitable type of structure in areas of mining subsidence, the
following is recommended:
Accessed by Brandon & Associates refreshed 170613 on 23 Oct 2019 [OBSOLESCENT] (Document currency not guaranteed when printed)

(a) The structure should be articulated or sufficiently flexible to ensure that mining
movements will least affect each member.
(b) The structure should remain serviceable during the passage of the subsidence wave,
but minor members and finishes, e.g., fascia panels, which can easily be repaired or
replaced, may be allowed to suffer some damage.
(c) Consideration at design stage should be given to the inclusion of details to facilitate
preventative and remedial works should the working of the underlying coal reserves
eventuate, e.g., removable deck joints and provision of jacking points.
(d) Bridge superstructure should be a simply supported beam and slab arrangement, with
low torsional properties.
(e) Substructures should not resist ground movements and piled foundations should be
avoided wherever possible.
(f) Buried structures should be divided into short lengths. Passive pressures should be
considered in the design of such structures.
This policy is based on many years of experience in the United Kingdom, as outlined by
Davies and Smith (1977) (Ref. 74), Jones and Spencer (1977) (Ref. 75) and Geddes (1977)
(Ref. 76).
Ideally, the structural type should be statically determinate with a simple and accessible
support system and adequate provision of joints for changes in dimension or tilt [see
Department of Main Roads, NSW (1981) (Ref. 77)].

C20 FORCES FROM BEARINGS


The value of the coefficient of friction depends largely on the maintenance condition of the
bearings. It should be noted that the friction coefficients for steel plate on steel plate and
steel roller and rocker bearings can vary considerably depending on the condition of the
bearings. Therefore, a stainless steel plate sliding on PTFE is generally a more suitable
bearing surface. Additional details of average values of the coefficient of friction for
bearings other than stainless steel on PTFE are given in Black (1972) (Ref. 78).
Friction coefficients for stainless steel on PTFE vary depending on the applied pressure, the
filling used in the PTFE, temperature, surface finish and the surface lubrication. Friction
coefficients for design are given in AS 5100.4, Bridge design—Bearings and deck joints,
and further information is available in Taylor (1972) (Ref. 79).

www.standards.org.au  Standards Australia


AS 5100.2 Supp 1—2007 64

It should be noted that, in some cases, the stiffness of the substructure may be sufficiently
low such that the frictional forces, indicated by the bearing characteristics, cannot be
mobilized.

C21 CONSTRUCTION FORCES AND EFFECTS


C21.1 General
Some bridge structures are inherently stable and may be constructed in a number of possible
ways. Other designs rely on a particular method of construction, and the construction stage
loads may be a critical design factor. Such restrictions should be made clear in the
drawings.
Particular attention to instability problems of the partially completed structure and in the
handling of slender members should be taken. The effects of normal construction tolerances
should be considered and allowed for where necessary.
If construction traffic, for example concrete trucks, cranes, and the like, will be allowed on
a partially completed structure, the resulting effects of braking forces, vibration and impact
Accessed by Brandon & Associates refreshed 170613 on 23 Oct 2019 [OBSOLESCENT] (Document currency not guaranteed when printed)

should be considered. Any requirements for temporary restraints or early fixing, or both, of
permanent restraints should be specified.
The stability of a bridge under flood or wind forces may depend to some extent on the dead
load of the superstructure. Failure may occur if flood or wind forces occur prior to the
erection of the superstructure.
C21.2 Temporary structures
(No Commentary)

C22 LOAD COMBINATIONS


C22.1 Classification of loads and load effects
Thermal effects have been incorporated as an independent classification, rather than
included with permanent or transient effects. This is due to the fact that, although they are a
variable effect, they are significant for a much higher proportion of the time in comparison
with other transient effects.
C22.2 Ultimate limit state load combinations
The definition of strength and stability limit state load is such that the chance of two or
more ultimate transient or thermal effects occurring simultaneously is so remote, that it may
be neglected. In many cases, a strength and stability load event physically excludes other
extreme conditions.
It may be possible for some ultimate transient effects to occur in conjunction with other
transient or thermal effects at a serviceability design level. Such load combinations should
only be considered where it is likely that they may occur. For example, if the strength and
stability wind load is depicted by strong wind conditions, the serviceability level for
thermal loadings may also be reached; however, if the ultimate wind load is cyclonic, it is
unlikely that thermal loads would reach a serviceability level at the same time.
The ultimate wind load is considered to occur while the structure is loaded by railway
traffic load but only as a serviceability vertical load. In cases where the vertical load
improves the stability of the structure, the effect of a reduced vertical load with full wind
effects should be considered.
For large structures, if measures are taken by the authority to ensure that trains will not be
operating on the structure in extreme wind, the ultimate limit state wind loads applying to
railway live load may be reduced.

 Standards Australia www.standards.org.au


65 AS 5100.2 Supp 1—2007

C22.3 Serviceability limit state load combinations


The basic load combination is defined in such a way that the design engineer should
determine and consider the serviceability load combinations that may realistically occur
during the life of a structure.
The factor (k) takes into account the improbability of more than one transient effect
reaching its serviceability design level at the same time, or occurring simultaneously with
full serviceability thermal effects. The values of k have been chosen to reduce the
magnitude of the secondary transient or thermal effects to a frequently occurring value.
C22.4 Design loads specific to an element
(No Commentary)

C23 ROAD SIGNS AND LIGHTING STRUCTURES


C23.1 General
(No Commentary)
Accessed by Brandon & Associates refreshed 170613 on 23 Oct 2019 [OBSOLESCENT] (Document currency not guaranteed when printed)

C23.2 Limit states


C23.2.1 Ultimate limit state
(No Commentary)
C23.2.2 Serviceability limit state
Vibration of wind-sensitive sign or lighting structures can occur as a result of gusting,
where it is generally of a transient nature, or vortex shedding where, under certain
conditions, it may be a sustained vibration. The consequences of vibration can be, for
example, either fatigue or failure of electrical components. Even from a visual point of
view, it is considered undesirable to observe structures such as light standards exhibiting
significant cyclic displacement in a normal breeze.
The designer should attempt to preclude resonance due to vortex shedding within normal or
common wind speeds up to and including the serviceability wind speed, as this can
constitute a failure to meet the serviceability limit state requirements. Resonance occurs
where the frequency of vortex shedding (ne ) equals a structure resonance frequency (n i).
The critical wind speed (V cr), is calculated as follows:
d ni
V cr = . . . C23.2.2
S
where
d = diameter of a circular member or the frontal width of a prismatic member
S = Strouhal number, which is a function of the cross-sectional shape
If V cr falls within the design wind speed range, then the design engineer should investigate
the consequences. The value of n i may be increased by increasing structural stiffness,
thereby increasing V cr to a value greater than the maximum design wind speed.
Alternatively, the reverse may be carried out, reducing Vcr to a speed low enough to produce
very small vibratory amplitudes. The problem is examined by Ulstrup (1978) (Ref. 80), the
AASHTO Standard Specification for Structural Support for Highway Signs (2003) (Ref. 81)
and the Canadian Highway Bridge Design Code (Ref. 7).

www.standards.org.au  Standards Australia


AS 5100.2 Supp 1—2007 66

Primarily, good engineering judgement combined with experience will play an important
role in preventing vibratory failures. Joints and welds designed to produce minimum stress
raising effects should reduce the danger of fatigue. The following empirical criteria, which
have resulted in satisfactory design, may also be used and should be applied as
serviceability limit state requirements:
(a) For overhead sign structures (span type) The midspan deflection under dead load
should not exceed d 2/122, where d is the sign depth.
(b) For light poles The maximum deflection of the luminaire under the serviceability
design wind loading should not exceed 5% of the luminaire mounting height.
C23.3 Design wind speeds
C23.3.1 Ultimate limit state
While the strength and stability limit state design wind load for bridge structures is based
on a 2000 year return interval, it is not considered necessary to apply such a stringent
condition to highway sign, lighting and signal structures.
Accessed by Brandon & Associates refreshed 170613 on 23 Oct 2019 [OBSOLESCENT] (Document currency not guaranteed when printed)

This is primarily because of the shorter design and useful life of such structures and the
generally less severe consequences of failure.
In the case of traffic sign structures and traffic signal mast arms, that span or cantilever
over traffic lanes, a 1000 year return interval has been chosen which corresponds
approximately to a 50 year design life. This is intended to reflect the relative difficulty of
replacing such structures and the potential risk of failure to road traffic. Reference may be
made to AS 2979, Traffic signal mast arms, for the design of traffic mast arms.
A 200 year return interval design wind speed has been chosen for light poles, signs and
signals over 5 m high, except for the structures referred in this Clause. The design wind
speed for such structures should generally be about 45 m/s but may be more for relatively
high levels of exposure.
In the case of minor sign and signal structures, the 35 m/s design wind speed is considered
adequate for the following reasons:
(a) At wind speeds above 35 m/s, flying debris may well destroy the sign facing. On the
other hand, if the failure mode of the sign supports, which are generally metal tubing,
at this wind speed is ductile, the sign may be undamaged and restoration should be a
simple operation.
(b) Supports designed for this wind speed would not, in general, constitute a hazard to
vehicles.
C23.3.2 Serviceability limit state
For structures required to be designed for a 20 year return period, the design wind speed is
to be determined in accordance with AS/NZS 1170.2 and will vary according to the height
and exposure of the structure.
For serviceability requirements for minor sign structures, an overall speed of 20 m/s is
prescribed and this value is generally related to a return interval of 1 to 2 years. The
serviceability requirement is intended to cover a range of wind speeds which occur
frequently. In certain circumstances, the authority may see fit to vary this speed.
C23.4 Design wind pressure
In calculating the peak response of a wind-sensitive structure to a design wind gust, it is
common to use Bernoulli’s equation, wherein the dynamic pressure is represented by a
static one. The wind speed used in this calculation is the peak, 3 s gust speed for the chosen
return interval.

 Standards Australia www.standards.org.au


67 AS 5100.2 Supp 1—2007

In structures that are relatively flexible, such as light poles, this method can be significantly
on the non-conservative side and research has shown that the peak response is more
accurately described in terms of the mean hourly wind speed and a gust or response factor.
This gust factor depends on the natural frequency of vibration of the structure, the height
above ground level and the amount of damping present. The gust energy or gust factor
method is discussed by Davenport (1967) (Ref. 82), Vickery (1971) (Ref. 83), the UK
Association of Public Lighting Engineers (Ref. 84) and AS/NZS 1170.2.
C23.5 Design loads
The design loads for these types of structures is invariably the combination of gravity
forces and wind load only.
C23.6 Service life loads on walkways
The additional load for walkways and service platforms is to provide for the stacking of
materials during construction and servicing.
Accessed by Brandon & Associates refreshed 170613 on 23 Oct 2019 [OBSOLESCENT] (Document currency not guaranteed when printed)

C24 NOISE BARRIERS


C24.1 Wind pressure on noise barriers
(No Commentary)
C24.2 Average return interval
The ultimate limit state average return interval of 2500 years is intended to reflect the
importance of ensuring that roads and railway lines critical to post-disaster recovery, from
events such as earthquakes and cyclones, are not blocked by failed noise walls.
The ultimate limit state average return interval of 200 years is intended to reflect the
relatively low risk to third parties and cost of repairs for noise walls located on property
owned by road or rail authorities.
The ultimate limit state average return interval of 500 years for the design of all other noise
walls reflects the intermediate level of risk compared to the situations described in this
Clause.
C24.3 Design life
(No Commentary)
C24.4 Change in terrain category
(No Commentary)
C24.5 Shielding multiplier (Ms)
(No Commentary)
C24.6 Topographic multiplier
(No Commentary)
C24.7 Net pressure for hoardings and freestanding walls
(No Commentary)
C24.8 Free ends
(No Commentary)

www.standards.org.au  Standards Australia


AS 5100.2 Supp 1—2007 68

APPENDIX CA
DESIGN LOADS FOR MEDIUM AND SPECIAL PERFORMANCE LEVEL
BARRIERS

CA1 GENERAL
(No Commentary)

CA2 DESIGN LOADS


CA2.1 Medium performance level barriers
For background information on these design loads, reference should be made to
Clause C11.2.2. The medium performance level design loads and contact lengths
correspond approximately to the AASHTO LRFD (2400) (Ref. 15) test level 5 design
Accessed by Brandon & Associates refreshed 170613 on 23 Oct 2019 [OBSOLESCENT] (Document currency not guaranteed when printed)

parameters.
CA2.2 Special performance level barriers
Design forces and contact lengths are provided for two specific special performance level
barriers. The first of these special performance levels corresponds to the AASHTO LRFD
(2004) (Ref. 15) test level 6. The second of these special performance levels corresponds
approximately to the AASHTO (1989) (Ref. 26) performance level PL-4. The design forces
specified for these performance levels are respectively approximately 50% and 100%
greater than those specified for medium performance level barriers to provide clearly
recognisable steps in containment capacity over those defined for lower performance levels.
Each successive performance level, with the exception of the special performance level
corresponding to test level 6, corresponds to a doubling of the design forces.
As limited tested systems equivalent to these performance levels are available at this stage,
significant reliance must be placed on these design parameters for the detailing of
appropriate barrier systems. It is anticipated that a number of systems will be tested
internationally for compliance at the test level 6 but few if any at higher levels which places
significant reliance on developments in computer simulation.
The requirements for other special performance level barriers need to be defined by the
relevant authority.

CA3 EFFECTIVE HEIGHTS


The medium performance level effective height of 1100 mm is based on the theoretical
requirement, the typical tyre height of a heavy vehicle and the successful record in the USA
of 1067 mm high concrete parapets redirecting medium sized commercial vehicles.
The special performance level effective height of 1400 mm is aimed at containing the tray
of heavy high centre of gravity van type semi-trailers. This is to prevent them from
overturning.
In order to prevent a special performance level heavy tanker type semi-trailer from
overturning, a greater effective height, in the order of 1700 mm to 2000 mm may be
required for a combination parapet to prevent overturning of such a vehicle. Hence, in the
case of a very high-risk situation, an individual benefit-cost assessment is still required.
In the case of other special performance level barriers, the appropriate effective height shall
be determined and specified by the authority.

 Standards Australia www.standards.org.au


69 AS 5100.2 Supp 1—2007

REFERENCES
1 Austroads Bridge Design Code, ABDC-1992 (HB 77.2—1996).
2 Gordon, R. and Boully, G., ‘Revised Bridge Design Live Loads, Austroads Bridge
Conference, 1997, Sydney, Australia.
3 Gordon, R. and Boully, G., ‘Bridges—Planning for the Future’, 5th International
Symposium on Heavy Vehicle Weights and Dimensions, Maroochydore, Queensland,
March 1998, Australia.
4 Heywood, R., Gordon, R. and Boully, G., ‘Australia’s Bridge Design Load Model,
Planning for an Efficient Road Transport Industry’, 5th International Bridge
Engineering Conference, Jnl of the Transportation Research Board, Florida,
April 2000, No. 1696, Vol 2.
5 Ontario Highway Bridge Design Code and Commentary, Ministry of Transportation
and Communications, 1983, Ontario.
6 Clough, R.W. and Penziem, J., ‘Dynamics of Structures’, McGraw Hill Kogakusha
Accessed by Brandon & Associates refreshed 170613 on 23 Oct 2019 [OBSOLESCENT] (Document currency not guaranteed when printed)

Ltd., Tokyo, 1975.


7 Canadian Highway Bridge Design Code and Commentary, CSA International,
December 2000, Ontario, Canada.
8 Heywood, R. and Boully, G., ‘Dynamic Response of Bridges to Heavy Vehicles’
Proceedings Austroads 4th Bridge Conference, Nov 2003, Adelaide, Australia.
9 Heywood, R., Dynamic Interaction of Vehicles and Bridges—Testing of Chiltern/
Beechworth Overpass, Austroads Project NT & E 9909, Aug 2000.
10 Heywood, R., Dynamic Interaction of Vehicles and Bridges—Review of Bridge Tests
& Recent Research, Austroads Project NT & E 9909, Aug 2000.
11 Green, R., ‘Dynamic Response of Bridge Superstructures—Ontario Observations’,
TRRL Symposium on Dynamic Behaviour of Bridges, Supplementary Report No. 257,
1977.
12 Prem, H. and Heywood, R., ‘Effects of Road Profile Rehabilitation on Dynamic
Loading of Bridges’, ARRB Transport Research, Vic Roads R&D Project 717, 2000.
13 Roberts, S. and Heywood, R., ‘Fatigue—It will Change our Culture’, 5th Austroads
Bridge Engineering Conference, 2004, Tasmania.
14 Grundy, P. and Boully, G., ‘Fatigue Design in the New Australian Bridge Design
Code’, 5th Austroads Bridge Engineering Conference, 2004, Tasmania.
15 AASHTO LRFD, ‘Bridge Design Specifications’, American Association of State
Highway and Transport Officials, 3rd Edition, Washington, D.C., 2004.
16 Marcer, J.R., ‘Revised Railway Provisions of the New Bridge Design Standard’,
Conference on Railway Engineering, November 2002, Wollongong.
17 UIC-776-1 R, ‘Loads to be considered in railway bridge design’, 4th Edition,
July 1994.
18 Tajadini, A., ‘Testing the Effects of Wheel Impact Loads on Bridges’, AAR Research,
Test and Development Report, Railway Track and Structures, March 1996.
19 Marcer, J.R., ‘Design Aspects Significantly Affecting Rating and Fatigue Life of
Railway Underbridges’, Asia Pacific Conference on Bridge Loading and Fatigue,
December 1996, Monash University.
20 American Railway Engineering Association (AREA), Chapter 15, Iron and Steel
Structures, 1.3.5 Impact (a)2, 1969.

www.standards.org.au  Standards Australia


AS 5100.2 Supp 1—2007 70

21 Fisher, W., ‘Bridge Fatigue Guide—Design and Details’, American Institute of Steel
Construction, 1977.
22 Report on the Formal Investigation of an Accident on or about the Main Western
Railway Line at Granville on 18th January 1997 by His Honour Judge J.H. Staunton,
Q C, Chief Judge of the District Court, 11th May 1977.
23 NAASRA, ‘Bridge Design Specification’, National Association of Australian State
Road Authorities, Sydney, 1976.
24 NCHRP Report 350, ‘Recommended Procedures for the Safety Performance
Evaluation of Highway Features’, Transportation Research Board, National Research
Council, Washington, D.C., 1993.
25 U.S. Department of Transportation, FHWA Memorandum to Regional Highway
Administrators, 1993.
26 AASHTO, ‘Guide Specification for Bridge Railings’, 1989.
27 AASHTO LRFD, ‘Bridge Design Specifications’, American Association of State
Accessed by Brandon & Associates refreshed 170613 on 23 Oct 2019 [OBSOLESCENT] (Document currency not guaranteed when printed)

Highway and Transportation Officials, 1st Edition, Washington, D.C., 1994.


28 BS 6779, ‘Highway Parapets for bridges and other structures, Part 1: Specifications
for vehicle containment parapets of metal construction’, ‘Part 2: ‘Specifications for
vehicle containment parapets of concrete construction’, London 1992.
29 EN 1317-2, ‘Road Restraint Systems—Part 2: Performance Classes, Impact Test
Acceptance Criteria and Test Methods for Safety Barriers’, July 1998.
30 Bronstad, M.E., Michie, J.D., ‘Multiple service level highway bridge railing selection
procedures’. NCHRP, TRB, Reports 239, Washington, November 1981.
31 Colosimo, V., ‘Design Guidelines For Bridge Barriers’, AUSTROADS Bridge
Conference, Bridging the Millenia, Sydney, December 1997.
32 Colosimo, V., ‘Selection Guidelines for Bridge Barriers, Road safety & Traffic
Enforcement’, ITE International Road Safety & Traffic Enforcement Conference,
Beyond 2000, 6–7 September 1999.
33 Colosimo, V., ‘Bridge Barriers—Standards and Code Provisions’, AUSTROADS 4th
Bridge Conference, Adelaide, November 2000.
34 Colosimo, V., ‘Bridge Barriers—Implementing the AS 5100 Bridge Design Code
Provisions’, AUSTROADS 5th Bridge Conference, Tasmania, May 2004.
35 Hirsch, T.J., ‘Analytical evaluation of Texas bridge rails to contain buses and trucks’.
Texas transportation Institute T.S.D.H & P.J., Research Report 230-2, August 1978.
36 Hirsch, T.J., ‘Longitudinal Barriers for buses and trucks’. Texas transportation
Institute. Reported in Transportation Research Record 1052, Symposium on
Geometric Design for Large Trucks, pp 95-102, TRB, NRC, 1986.
37 Troutbeck, Prof., R.T. and Bunker, J.B., ‘Rational Standards for Bridge Barriers,
procedure for Estimating Collision Frequency and Severity for Bridge Barriers’,
Queensland University of Technology (QUT), 1994.
38 Wheeler, J.E., Prediction and Control of Pedestrian Induced Vibrations in
Footbridges, Journal of the Structural Division, ASCE, Vol 108, No. ST9,
September 1982.
39 Irwin, A.W., Human Response to Dynamic Motion of Structures, Structural Engineer,
Vol 56A No. 9, September 1978.

 Standards Australia www.standards.org.au


71 AS 5100.2 Supp 1—2007

40 ACI 215-74, Considerations for the Design of Concrete Structures Subjected to


Fatigue Loadings, Journal of the American Concrete Institute, Proc. Vol. 71, No. 3,
March 1974.
41 BS 5400-2, ‘Steel, concrete and composite bridges—Specification for loads’, 1978.
42 Wright, R.N. and Walker, W.H., ‘Vibration and Deflection of Steel Bridges’, AISC
Engineering Journal, May 1971.
43 Wheeler, J.E., Bridge Dynamics in a Limit States Concept, Main Roads Department
of Western Australia, Technical Report No. 24, September 1981.
44 Wheeler, J.E., Pedestrian-Induced Vibrations in Footbridges, Xth ARRB Conference,
Sydney 1980.
45 Dallard et al, ‘The London Millennium Footbridge’, The Structural Engineer, 20 Nov
2001.
46 Dallard, P., ‘London Millenium Bridge: Pedestrian-Induced Lateral Vibration’,
Journal of Bridge Engineering (ASCE), No. 6, Vol 6 Nov-Dec 2001.
Accessed by Brandon & Associates refreshed 170613 on 23 Oct 2019 [OBSOLESCENT] (Document currency not guaranteed when printed)

47 Teng, W.C., ‘Foundation Design’, Prentice-Hall, New Jersey, 1962.


48 Terzaghi, K., ‘Anchored Bulkheads’, Transactions ASCE, CXIX, 1954.
49 AREA Manual, American Railway Engineering Association, ‘Part 5: Retaining Walls
& Abutments’, 1989 (Clause C13.3).
50 Bull, K.D., ‘Seismic Design of Highway Structures’, Journal of Structural Division,
ASCE, ST8, August 1972.
51 Newmark, N.M. and Rosenbleuth, E., ‘Fundamentals of Earthquake Engineering’,
Prentice-Hall Inc., London, 1971.
52 New Zealand National Society for Earthquake Engineering Bulletin, ‘Seismic Design
of Bridges’, 1980. (Available as New Zealand National Roads Board, Road Research
Bulletin 56, 1981).
53 Japanese Society of Civil Engineers, ‘Earthquake Resistant Design for Civil
Engineering Structures, Earth Structures and Foundation in Japan’, 1973.
54 Berrill, J.B. and Elms, D.G., ‘Seismic Response of Retaining Walls and a Brief
Survey of Soil-Structure Interaction’, Road Research Unit, National Roads Board,
New Zealand, Bridge Seminar, Auckland, 1978.
55 Ferriott, J.M. and Forrest, J.B., ‘Determination of Seismically Induced Soil
Liquefaction Potential at Proposed Bridge Sites’, FHWA Report FHWA-RD-77-127/8,
1977.
56 Seed, H.B. and Whitman, R.V., ‘Design of Earth Retaining Structures for Dynamic
Loads’, ASCE Specialty Conference—‘Lateral Stresses in the Ground and Design of
Earth Retaining Structures’, ASCE, New York, pp 103–147, 1970.
57 Richards, R. and Elms, D.G., ‘Seismic Behaviour of Gravity Retaining Walls’,
Journal of Geotechnical Division, ASCE, Vol 105, No. GT4, pp 449–464, 1979.
58 Park, R. and Blakeley, R.W.G., ‘Seismic Design of Bridges’, Road Research Unit,
National Roads Board, New Zealand, Bridge Seminar, Auckland, 1978.
59 New Zealand Ministry of Works & Development, Highway Bridge Design Brief,
July 1973.
60 Hogben, N., Miller, B.L., Searle, J.W. and Ward, G., ‘Estimation of Fluid Loading on
Offshore Structures’, Proceedings, Institution of Civil Engineers, Part 2, Vol 63,
September 1977.

www.standards.org.au  Standards Australia


AS 5100.2 Supp 1—2007 72

61 Leonard, J.W., Garrison, C.J. and Hudopeth, R.T., ‘Deterministic Fluid Force on
Structures’, A Review, Journal of Structural Division, Proceedings, ASCE, ST6,
June 1981.
62 Apelt, C.J., ‘Flow Loads on Bridge Piers’, Journal of Institution Engineers Australia,
July/August 1965.
63 Apelt, C.J. and Isaacs, L.T., ‘Bridge Piers, Hydrodynamic Force Coefficients’,
Journal of Hydraulics Division, Proceedings, ASCE, January 1968.
64 Jempson, M.A. and Apelt, C.J., ‘Flood loads on bridge superstructures, Bridges into
the 21st Century, Hong Kong’, The Hong Kong Inst. of Engineers, Hong Kong,
1025–1032, 2–5 October 1995.
65 Jempson, M.A. and Apelt, C.J., ‘Flood loads on submerged and semi-submerged
bridge superstructures, Bridging the Millenia’, Austroads 1997 Bridge Conference,
‘Bridging the Millenia’, Sydney, 3–5 December 1997, ed Chirgwin, G.J., Austroads
Inc., Sydney, Australia, 2, 19–33.
Accessed by Brandon & Associates refreshed 170613 on 23 Oct 2019 [OBSOLESCENT] (Document currency not guaranteed when printed)

66 Jempson, M.A. and Apelt, C.J., ‘Debris loadings on bridge superstructures and piers,
Bridging the Millenia’, Austroads 1997 Bridge Conference, ‘Bridging the Millenia’,
Sydney, 3–5 December 1997, ed Chirgwin, G.J., Austroads Inc., Sydney, Australia,
2, 3–17.
67 Jempson, M.A., ‘Hydrodynamic forces on partially and fully submerged bridge
superstructures’, M Eng Sc Thesis, University of Queensland, Australia, 1995.
68 Jempson, M.A., ‘Flood and debris loads on bridges’, PhD Thesis, University of
Queensland, Australia, 2000.
69 Thomas, J.D., ‘Temperature Effects on Bridge Structures’, CRB—University of
Melbourne Research Project (Commonwealth Project No. 75/342), Bridge Design
Methods Research, October 1980.
70 Priestly, M.J.N. and Buckle, I.G., ‘Ambient Thermal Response of Concrete Bridges’,
RRU Bulletin 42, National Roads Board, Wellington, New Zealand, 1979.
71 National Coal Board, Production Department, Subsidence Engineers Handbook,
London, 1975.
72 Frankham, B.S. and Mould, G.R., ‘Mining Subsidence in NSW Recent Developments,
Australasian Institute of Mining and Metallurgy Conference, New Zealand,
May 1980.
73 Sims, F.A. and Bridle, R.J., ‘Bridge Design in Areas of Mining Subsidence, The
Journal of the Institution of Highway Engineers, November 1966.
74 Davies, B.L. and Smith, R., ‘The Influence of Coal Mining on the Maintenance,
Design and Construction of Highway Bridges and County-Owned Structures in South
Yorkshire’, Proceedings of the Conference on Large Ground Movements and
Structures, UWIST, Cardiff, 1977.
75 Jones, C.J. and Spencer, W.J., ‘The Implications of Mining Subsidence for Modern
Highway Structures’, Proceedings of the Conference on Large Ground Movements
and Structures, UWIST, Cardiff, 1977.
76 Geddes, J.D., ‘Construction in Areas of Large Ground Movement’, Proceedings of the
Conference on Large Ground Movements and Structures, UWIST, Cardiff, 1977.
77 Department of Main Roads, N.S.W., ‘Bridge Structures in Areas of Mining
Subsidence’, October 1981.
78 Black, W., ‘Notes on Bridge Bearings’, RRL Report LR 382, Crowthorne, 1972.

 Standards Australia www.standards.org.au


73 AS 5100.2 Supp 1—2007

79 Taylor, M.E., ‘PTFE in Highway Bridge Bearings’, RRL Report LR 491, Crowthorne,
1972.
80 Ulstrup, C.C., ‘Natural Frequencies of Axially Loaded Bridge Members, Proceedings,
ASEC, Journal of Structural Division, ST2, February 1978.
81 AASHTO, ‘Standard Specification for Structural Supports for Highway Signs,
Luminaires and Traffic Signals’, Article 1.9.6, 4th Edition, 2003.
82 Davenport, A.G., ‘Gust Loading Factors’, Journal of Structural Division, ASCE,
Vol 93, 1967.
83 Vickery, B.J., ‘On the Reliability of Gust Loading Factors’, Civil Engineering
Transactions, Institution of Engineers Australia, Vol 13, 1971.
84 U.K. Association of Public Lighting Engineers High Mast Lighting, Technical Report
No. 7, 1976.
Accessed by Brandon & Associates refreshed 170613 on 23 Oct 2019 [OBSOLESCENT] (Document currency not guaranteed when printed)

www.standards.org.au  Standards Australia


Accessed by Brandon & Associates refreshed 170613 on 23 Oct 2019 [OBSOLESCENT] (Document currency not guaranteed when printed)

AS 5100.2 Supp 1—2007


74

NOTES
Accessed by Brandon & Associates refreshed 170613 on 23 Oct 2019 [OBSOLESCENT] (Document currency not guaranteed when printed)

75

NOTES
AS 5100.2 Supp 1—2007
Accessed by Brandon & Associates refreshed 170613 on 23 Oct 2019 [OBSOLESCENT] (Document currency not guaranteed when printed)

AS 5100.2 Supp 1—2007


76

NOTES
Standards Australia
Standards Australia develops Australian Standards® and other documents of public benefit and national interest.
These Standards are developed through an open process of consultation and consensus, in which all interested
parties are invited to participate. Through a Memorandum of Understanding with the Commonwealth Government,
Accessed by Brandon & Associates refreshed 170613 on 23 Oct 2019 [OBSOLESCENT] (Document currency not guaranteed when printed)

Standards Australia is recognized as Australia’s peak non-government national standards body. Standards Australia
also supports excellence in design and innovation through the Australian Design Awards.

For further information visit www.standards.org.au

Australian Standards®
Standards®
Committees of experts from industry, governments, consumers and other relevant sectors prepare Australian
Standards. The requirements or recommendations contained in published Standards are a consensus of the views
of representative interests and also take account of comments received from other sources. They reflect the latest
scientific and industry experience. Australian Standards are kept under continuous review after publication and are
updated regularly to take account of changing technology.

International Involvement
Standards Australia is responsible for ensuring the Australian viewpoint is considered in the formulation of
International Standards and that the latest international experience is incorporated in national Standards. This role is
vital in assisting local industry to compete in international markets. Standards Australia represents Australia at both
the International Organization for Standardization (ISO) and the International Electrotechnical Commission (IEC).

Sales and Distribution


Australian Standards®, Handbooks and other documents developed by Standards Australia are printed and
distributed under license by SAI Global Limited.
Accessed by Brandon & Associates refreshed 170613 on 23 Oct 2019 [OBSOLESCENT] (Document currency not guaranteed when printed)

For information regarding the development of Standards contact:


Standards Australia Limited
GPO Box 476
Sydney NSW 2001
Phone: 02 8206 6000
Fax: 02 8206 6001
Email: mail@standards.org.au
Internet: www.standards.org.au

For information regarding the sale and distribution of Standards contact:


SAI Global Limited
Phone: 13 12 42
Fax: 1300 65 49 49
Email: sales@sai-global.com

ISBN 0 7337 8062 8


Accessed by Brandon & Associates refreshed 170613 on 23 Oct 2019 [OBSOLESCENT] (Document currency not guaranteed when printed)

This page has been left intentionally blank.

You might also like