You are on page 1of 74

Bridge Launching

ISBN 978-0-7277-5997-9

ICE Publishing: All rights reserved


http://dx.doi.org/10.1680/bl.59979.225

Chapter 4
Steel and composite bridges

In recent decades, many factors have increased the competitiveness of composite bridges and
reduced their lower threshold of utilisation to spans of about 50 m, which previously were the
dominion of prestressed concrete (PC) box girders.

Progress in iron metallurgy led to rolled steel of high and reliable mechanical properties. The
development of design standards based on the strength of materials led to a better identication
of structural safety. Progress in assembly techniques replaced riveted joints with bolted joints rst,
and eventually with welded joints. The composite action of concrete and steel has been thoroughly
investigated, and comprehensive design approaches have been ratied in the design standards.
Finally, over the past few years, this evolution has been further accelerated by the stagnant price
of steel plates and a general increase in labour costs (Matildi and Matildi, 1990).

Composite bridges offer several advantages. The high tensile and shear strength of the steel plates,
combined with the low-cost compressive strength of concrete, leads to light and efcient
structures. Rapidity of construction generates nancial savings, building most of the bridge in a
workshop enhances quality, and small dimensional tolerances enhance the reliability of structural
analysis. Renewable protective treatments offer long-term durability, and the possibility of
replacing deteriorated components and adapting the structure to new work conditions offers
design exibility. Demolition is less critical than for PC bridges, and recycling most of materials
further reduces the total cost. Finally, the architectural quality of structural elements whose
function is clearly recognisable enhances bridge aesthetics.

The different efciency of PC and composite bridges derives from the different efciencies of
materials. The efciency of a material may be expressed by the ratio of strength to specic weight
(Petrangeli, 1993), em = f/g. A concrete with compressive strength fc = 45 MPa and specic
weight gc = 25 kN/m3 has compressive efciency em 1.8 103 m and lower tensile and shear
efciency. An Fe510 EN 10025 steel plate with fy = 355 MPa and gs = 77 kN/m3 has tensile
efciency em 4.6 103 m at rst yielding and higher ultimate tensile strength, and is therefore
2.5 times more efcient at rst yielding. Although buckling reduces the compressive efciency
of a steel plate, the steel and composite bridges are more efcient than the PC ones, as their masses
are more distant from the gravity axis and the materials perform better.

Composite sections are particularly suited for simply supported beams. In a continuous beam, the
positive bending regions take full advantage of the composite action, while the negative bending
regions are problematic, both for the steel girders and for the concrete slab. High compressive
stresses in the bottom anges require thick plates or closely spaced stiffeners to prevent buckling,
a cracked slab contributes to the composite moment of inertia less effectively, and large quantities
of reinforcement are necessary in the concrete slab to control crack width and ensure durability.

225

Downloaded by [] on [25/09/17]. Copyright ICE Publishing, all rights reserved.


Bridge Launching

Advances in metallurgy and workshop technology also led to the incremental launching construc-
tion of steel bridges with an orthotropic deck plate. Because of the higher construction cost, these
bridges are typically launched on spans longer than those achievable with composite cross-
sections. Several long-span steel bridges have been built by full-span launching or with the help
of temporary piers. The Millau Viaduct in France is probably the most famous of the long-span
launched bridges.

4.1. Conceptual design and pre-sizing


For cost and maintenance reasons, most steel girders erected by incremental launching have a
concrete slab. Design of the steel girder must be compatible with the transient launch stresses,
which mostly depend on self-weight. The concrete slab is 7585% of the total weight of the
composite cross-section and, although the slab provides substantial stiffness, the steel girders are
launched alone and set on the permanent bearings, and the deck slab is then cast in-place with
forming carriages or is assembled over the steel girders (Rosignoli, 2013).

Exceptions are always possible. A 675 m highway bridge was built in Germany by incrementally
launching the nished composite deck over temporary piers (Saul, 1997). Two 194 m composite
bridges were launched full-span in the UK with the use of thicker webs to control local buckling
during launching. Launching the nished girder is also standard practice when the girder has a
steel orthotropic deck.

When the steel girder is launched alone and the concrete slab is cast later, the girder is designed to
resist the weight of slab and forming systems prior to the onset of composite action. Since this load
is 34 times greater than the weight of the steel girder, huge margins of strength seem available for
the launch stresses. In reality, composite bridges are competitive for spans longer than 50 m, and
these spans suggest the use of steel girders with constant depth but varying cross-section. Flange
dimensions and web thickness vary along the girder according to the exural and shear demand of
the composite deck and to technological and economic requirements. Casting the concrete slab
raises the neutral axis of the composite section, which governs the nal distribution of exural
stresses, and provides tributary area and lateral stability to the top ange. For all these reasons,
the launch may be demanding even if the steel girder is lighter than the nal composite bridge.
Local stress concentration further complicates the situation.

The launch stresses rarely govern pre-sizing of the steel girder, which is carried out with conven-
tional criteria for composite bridges. The launch stresses are controlled with the launch devices
and, if necessary, by modifying some details of the girder design. The design of the cross-
section depends on the concrete-to-steel ratio of the composite section, on the weight-to-cost
ratio of its components, on the fabrication and assembly costs of the steel girder, on the
construction cost of the concrete slab, on the desired slenderness and on the need for torsional
stiffness.

As the concrete slab is a major contributor to the weight of the cross-section, its thickness is a
prime design parameter. The slab thickness depends on the deck breadth, the transverse spacing
and relative stiffness of girders and substringers, the strength of concrete, the presence of
transverse post-tensioning, and the required cover to top and bottom reinforcement. If shear
connectors are provided throughout the length of the bridge, the concrete slab cooperates with the
steel girder in resisting exural stresses, and cross-sections under positive and negative bending are
considered composite.

226

Downloaded by [] on [25/09/17]. Copyright ICE Publishing, all rights reserved.


Steel and composite bridges

When compressed, the concrete slab is often considered as fully cooperating for structural
analysis. Most design standards specify the use of a reduced effective breadth of concrete slab for
the cross-sectional checks to account for shear-lag effects in thin slabs. Some design standards also
specify the use of effective properties for structural analysis.

4.1.1 Multi-girder systems


The cross-section of the rst composite bridges contained several trusses or I-girders that
supported a concrete slab. The idea behind the structural concept was that multiple I-girders
provide alternative load paths in the case of failure of a girder. Other advantages derive from the
short spacing of the girders and the narrow side wings of the concrete slab, which allow the use of
200250 mm precast deck panels. Finally, multiple I-girders require thin ange plates and simple
welds at the changes of ange thickness.

Today, multi-girder decks are the standard solution only in the USA (Figure 4.1). Up to the 1970s,
many bridges were also designed with twin I-girders in the USA, sometimes combined with oor-
beams or substringers (Azizinamini, 2013). The perceived notion that twin I-girder decks are not
redundant led to a signicant reduction in their use, even though the commercial availability of
high-grade steels with high fracture toughness has recently led to a re-evaluation of twin I-girder
decks.

Multiple I-girders have a higher quantity of eld splices and cross-frames (Figure 4.2), while the
low cost of steel plates and the ever-increasing cost of labour suggest the minimisation of work-
shop and eld assembly activities. When the girders are erected with ground cranes, slow erection
combines with the risks of massive air working. Painting and maintenance costs increase linearly

Figure 4.1 Launch of the Clifford Hollows Bridge

227

Downloaded by [] on [25/09/17]. Copyright ICE Publishing, all rights reserved.


Bridge Launching

Figure 4.2 Launch of adjacent multi-girder systems

with the number of girders. Thin concrete slabs have caused durability issues due to fatigue and
corrosion of reinforcement in the past. A thicker concrete cover to top reinforcement reduces the
exural efciency of reinforcement and contradicts the need for controlling crack width for
enhanced protection of reinforcement.

The divergence between the cost of labour and steel plates has led to structural solutions for com-
posite bridges that are simpler to fabricate, erect and maintain. Trusses have been abandoned to
the advantage of I-girders, and multi-girder systems are progressively being abandoned to the
advantage of girdersubstringer systems, twin I-girders, and U-girders for composite box sections.
Twin I-girders offer similar levels of robustness and redundancy as the multi-girder systems,
because stiffer bracing and a thicker concrete slab provide alternative load paths in the case of
failure of one girder. The girders are assembled on the ground and erected in one lift, and air work
is limited to two eld splices per span (Bernard, 1997). The concrete slab is typically cast in-place
after completion of girder erection, but full-depth precasting of deck panels may also be adopted in
cold weather and for accelerated minimal impact construction.

In spite of their weak points, multi-girder systems are still of interest for wide decks and slender
bridges (Hw  L/30 where Hw is the total depth of the steel girder), or to facilitate crane assembly
of paired girders. They are also of interest on short spans because of the possibility of using less
expensive commercial rolled beams.

4.1.2 Girdersubstringer systems


Built-up I-girders distributed with 67 m spacing are the main load-carrying members of a girder
substringer system. The girders are connected by inverted K-frames that support rolled substringers

228

Downloaded by [] on [25/09/17]. Copyright ICE Publishing, all rights reserved.


Steel and composite bridges

with 56 m longitudinal spacing. The substringers are made with 600700 mm commercial rolled
shapes that are less expensive than built-up I-girders and provide additional support lines for the
concrete slab. The 33.5 m transverse spans of the deck slab allow the use of 200250 mm full-depth
precast panels. Compared with a multi-girder framing system, a smaller number of bearings and
cross-frames and 1020% savings in steel weight are additional advantages (Rosignoli, 2014).

Girdersubstringer systems are typically used on spans longer than 80100 m. Wide decks may
require ve main girders and four substringers; this solution is being adopted for the 29.3 m-wide
approaches of the New NY Bridge in the USA to support 250 mm full-depth precast deck panels.
The 20.7 m-wide deck of the Park Bridge in Canada (Gale, 2011) required four main girders and
three substringers to support a 230 mm two-layer slab comprising 110 mm precast panels and a
120 mm in-place layer.

Girdersubstringer systems are easier to launch than multi-girder framing systems because of the
smaller number of support lines for the deck. When the deck has four I-girders, two pairs of
braced girders may be launched independently (Figure 4.3). When the deck has ve I-girders, a
two-girder assembly and a three-girder assembly are launched. On launch completion, the two
assemblies are connected with the middle bay of K-frames and substringers. Bolted eld splices
between substringers and K-frames avoid fatigue due to the different longitudinal stiffness of
substringers and main girders.

This framing system is often combined with the use of full-depth precast slab panels or two-layer
decking systems including self-supporting precast panels and an in-place layer. Thin left-in-place

Figure 4.3 Launch of the Park Bridge girdersubstringer system. (Photo: Somerset/KWH)

229

Downloaded by [] on [25/09/17]. Copyright ICE Publishing, all rights reserved.


Bridge Launching

Figure 4.4 Top and bottom lateral braces in the first two spans. (Photo: Somerset/KWH)

concrete panels are rarely considered in the analysis of the exural capacity of the composite cross-
section and of the local punching shear and exural capacity of the concrete slab. Precast planks
for 3 m transverse spans are 4060 mm thick and are used only on short bridge spans because of
weight reasons.

One plane of lateral braces between the main I-girders is indispensable throughout the length of
the bridge for launch stability and to resist lateral wind prior to casting of the concrete slab. The
lateral braces are applied at the bottom ange level to provide torsional stiffness after completion
of the composite section and to help maintain the main girders in position over the launch rollers;
the launch systems are therefore designed to avoid interference. Curved bridges and long-span
launches suggest the use of top and bottom lateral braces in the rst two spans for control of lateral
torsionexure instability of the front cantilever in combination with the K-frames (Figure 4.4).

Two levels of lateral bracing and heavier anges and webs in the front deck region increase the
weight of the steel girder. For the full-span launch of the curved girdersubstringer system of the
Park Bridge in Canada, on 80 m spans, the increase in weight was around 7% (Gale, 2011).

4.1.3 Twin I-girders


The advances in welding techniques of thick plates permitted the development of composite
bridges based on two braced I-girders (Figure 4.5). The twin I-girder decks range from 50 m spans
to the much longer spans of cable-stayed bridges. Compared with the multi-girder systems, the

230

Downloaded by [] on [25/09/17]. Copyright ICE Publishing, all rights reserved.


Steel and composite bridges

Figure 4.5 Twin I-girder system

twin I-girder decks require less web stiffeners, eld splices, cross-frames and lateral bracing, and
less labour for fabrication and site assembly.

When the deck is 1520 m wide, the concrete slab may be supported on the main girders and on
oorbeams spaced 45 m, as in Figure 2.40. This solution allows for a constant 0.20.25 m slab
thickness, which facilitates full-depth precasting. Precast deck panels are connected to main girders
and oorbeams by lling continuous joints and stud pockets over the anges with non-shrink
concrete or special grouts. This structural scheme is frequently used in cable-stayed bridges, where
the main girders are located at the edges of the deck to anchor the stay cables with anchor pipes
bolted to the outer face of the webs or extending over them.

When the deck is narrower, the concrete slab is supported only on the main girders, which are
spaced about 55% of the deck width and braced by cross-frames or mid-depth cross-beams framed
to web stiffeners. A reasonable thickness of the concrete slab (0.25 m in midspan and 0.30 m above
the girders) permits reaching 7 m spacing of the main girders and 13 m deck width. The depth of
the I-girders may be pre-sized as Hw = L/28.

The concrete slab is heavier than in multi-girder systems because of fewer support lines, and this can
be a limit on the longest spans. Cross-frames and cross-beams are spaced around 8 m longitudinally
to reduce their cost. The concrete slab is typically cast in-place segmentally on left-in-place forms
or with self-repositioning forming carriages carried by the main girders (Figure 4.6). Full-depth
precasting of the deck panels is also frequent in cold weather applications.

Some design standards restrict the longitudinal spacing of the cross-frames for curved bridges.
These limitations may be insufcient to ensure stability during launching, as the cross-frames
should be closer the shorter the plan radius. In order to reduce the warping stress of non-uniform
torsion, the spacing of cross-frames should not exceed 4.5 m for plan radii smaller than 60 m,
5.0 m for 60150 m radii, 6.0 m for 150300 m radii and 7.5 m for larger radii.

Cross-frames and pier diaphragms are primary load-carrying members in a twin I-girder deck.
Their function is to resist torsion and distortion and to provide lateral support to the compression
anges, and therefore they are attached to the main girders close to the anges. This is especially
necessary in curved bridges; in rectilinear bridges, cross-beams located at the mid-plane of the
main girders may be connected to vertical web stiffeners designed for transverse frame action in
exure and shear (Figure 4.5). This scheme simplies the eld splices, and the cross-beams support
the bottom form table of the forming carriage for in-place casting of the concrete slab.

The concrete slab provides lateral stiffness and support to the compression anges. Cross-frames
and cross-beams are designed for the erection loads and to provide lateral support to the bottom

231

Downloaded by [] on [25/09/17]. Copyright ICE Publishing, all rights reserved.


Bridge Launching

Figure 4.6 Self-positioning forming carriage. (Photo: Doka)

anges in the negative bending regions. Lateral bracing between bottom anges is designed for
lateral wind and stability of the front cantilever during launching, and for torsion in service.

Negative bending in the support regions is typically larger than positive bending at the midspan. If
the concrete slab is not compressed articially, the longitudinal tensile stress often exceeds the ten-
sile strength of concrete, and the exural capacity of the composite section, the moment of inertia
for deection calculations, and stiffness factors for analysis of the continuous beam are calculated
considering the longitudinal reinforcement of the concrete slab to act compositely with the main
girders. Cost-effective combinations of ange plates and slab reinforcement may be developed by
designing the top anges for the construction stresses related to the launch of the girder and seg-
mental casting of the slab, and by adding reinforcement in the concrete slab to meet the ultimate
limit state (ULS) negative bending demand and to control the crack width under serviceability
limit state (SLS) load combinations.

Long-span twin I-girders may be designed for double composite action. A concrete slab is cast
between the bottom anges to close the cross-section in the negative bending regions. An open
section may be used in the midspan, or the bottom slab may be extended throughout the length
of the bridge to increase the torsional stiffness and exural capacity. The dual-track, 1209 m-long
Arroyo Las Piedras high-speed railway bridge in Spain was launched from the opposite abutments
on 63.5 m spans complete with precast panels for both concrete slabs, which were connected to
the I-girders on launch completion. The slab panels were applied to the front spans on launch
completion to use the light I-girders as a launch nose (Mato and Santos, 2007).

232

Downloaded by [] on [25/09/17]. Copyright ICE Publishing, all rights reserved.


Steel and composite bridges

Under positive bending and single composite action, the neutral axis of the cross-section is further
from the top ange the longer the span. The compression slab increases the composite moment of
inertia, but also increases the distance of the bottom tension ange from the neutral axis, and the
section modulus for the lower edge increases moderately. For an efcient composite action at the
midspan, the area of the bottom ange should be larger than the area of the top compression
ange. This would also lower the neutral axis of the composite section toward the centre of gravity
of the I-girders. Span lengths suitable for incremental launching construction require the use of
built-up girders, and the size and thickness of the top and bottom anges can be specied freely.

In the past, the exural capacity of the I-girders was adapted to the demand by adding cover plates
to the base anges. Today, webs and anges are obtained from plates directly supplied with the
desired dimensions. Thick plates with certied mechanical properties are readily available, and
permit reductions in the number of cover plates or their total avoidance by using ange plates with
varying thickness. Varying-thickness plates range between 20 and 150 mm, the variation rate in
thickness is usually 57 mm/m and the maximum thickness variation is about 40 mm. Interest
in varying-thickness plates has increased considerably in recent years, as they permit savings in
material and welding costs. Full-penetration butt joint welds can be located in less stressed areas,
or avoided entirely when changes in thickness occur at the eld splices. The use of varying-
thickness bottom anges with ush bottom surfaces also results in a regular launch surface.

The use of wide, thin ange plates is justied when the ange area is adjusted by varying the
plate width rather than the thickness. The use of ange plates of constant breadth and varying
thickness is often preferred in launched bridges, since it simplies the forming systems for in-
place casting of the concrete slab, the precast deck panels, and the lateral launch guides for the
bottom anges.

To avoid instability, the width-to-thickness ratio of compression anges must not exceed limits
specied by the design standards that allow for geometry imperfections and residual weld stresses.
The geometry requirements for a rectilinear girder are often expressed as

b Es
kcr (4.1)
t fy

where t is the plate thickness, Es is the modulus of elasticity of steel, fy is the characteristic yield
strength, and the buckling coefcient kcr and the plate width b are specied for different members
of the cross-section. When the main girders are erected with ground cranes, these geometry restric-
tions affect the bottom anges in the negative bending regions and the top anges in the midspan.
If the girder is erected by incremental launching, these restrictions affect both anges throughout
the length of the bridge.

These restrictions apply for rectilinear girders, whose anges are subject to a uniform compressive
stress. In a curved girder, the longitudinal axial stress in the anges varies across the plate width
because of the restrained warping of non-uniform torsion, yielding starts at the ange tip, and this
causes a sudden decrease in the buckling factor. Some design standards reduce the maximum
breadth-to-thickness ratio or the SLS compressive stress limit in the anges based on the length
of the unsupported ange between cross-frames, the radius of the plan curvature, the radius of
gyration and the width of the compression ange, and the ratio of the ange tip stress to the
uniform bending stress at the diaphragm location.

233

Downloaded by [] on [25/09/17]. Copyright ICE Publishing, all rights reserved.


Bridge Launching

According to Equation 4.1, the width of a compression ange with a 355 MPa yield point cannot
exceed 21.6 times its thickness. Some design standards specify additional geometry restrictions
related to the distance between the self-weight counter-exure points of the continuous beam with
equations for the effective ange width to be used for calculations. The dimensions of the ange
plates also depend on technological requirements such as launch bearings and guides, and
distribution and type of shear connectors for the top ange. As a result of these requirements, the
width of the top ange usually varies between 400 and 800 mm, and the width of the bottom ange
varies from 500 to 1200 mm when the design span varies from 30 to 100 m.

The web plates are typically thicker than the minimum shear strength requirements to facilitate
handling during fabrication and to limit the number of vertical stiffeners. The plate thickness is
rarely less than 12 mm so that the web strength is affected minimally by corrosion. The current
design trend is to limit the stiffening of web panels to vertical stiffeners and to take the post-
buckling strength into account in the design. Slender cross-sections such as Class 4 sections as per
Eurocode 3 (EN 1993-1-1:2005: UNI, 2005a) are often used in the support regions of medium- and
long-span composite decks, and bending redistribution due to slab cracking and web buckling at
the ULS are taken into account.

Eurocode 4 (EN 1994-2:2005: UNI, 2006) provides instructions to account for bending redistribu-
tion due to cracking of the concrete slab, while the effects of web buckling are considered for
cross-section assessment, but not for structural analysis. In reality, twin I-girder bridges with
Class 4 sections at the supports will likely have the webs in the post-buckling domain at the
ULS, and bending redistribution toward more rigid regions should be expected.

Bending redistribution depends on plastic rotations at the support sections of the continuous
beam, and a certain amount of ductility is therefore needed. As beam ductility increases, so does
the bending redistribution that can be achieved, up to the level of redistribution that corresponds
to the formation of a plastic mechanism. While the exural capacity of a cross-section decreases
gently when increasing the web slenderness, the rotation capacity diminishes abruptly, and very
slender sections rarely have satisfactory ductile behaviour.

The ratio of the net depth of the web plate hw (clear between anges) to the thickness tw must avoid
web panel instability under the thrust of the compression ange. Several design standards cover
this mode of instability. According to Eurocode 3 (EN 1993-1-1:2005: UNI, 2005a), it must be

Aw Es
0.55
hw Af f y
 (4.2)
tw h E
1+ w s
3rfy

where Aw and Af are the cross-sectional areas of the web and the compression ange, respectively,
and r is the radius of the plan curvature of the girder web. In a rectilinear girder with a 355 MPa
yield point and, say, Af = 2Aw, the limit ratio is about 224. In practice, it rarely exceeds 200. The
American Association of State Highway and Transportation Ofcials (AASHTO, 2003) species
an upper limit of 170 for curved bridges with vertically stiffened web panels, and similar equations
govern the application of longitudinal stiffeners. RPX-95 (1996) species an upper limit of 160 for
vertically stiffened webs in the support regions of the continuous beam, and 240 for the midspan

234

Downloaded by [] on [25/09/17]. Copyright ICE Publishing, all rights reserved.


Steel and composite bridges

sections. These limits increase to 250 and 350, respectively, in the presence of longitudinal
stiffeners.

The cross-beams are often built-up I-girders. They are spaced 45 m when directly supporting the
concrete slab, and about 8 m when they act as transverse frames. The web thickness is 1012 mm.
Flanges of transverse frames are 220300 mm wide and about 20 mm thick. When the oorbeams
directly support the concrete slab, the top ange is about 300 mm wide, to facilitate the application
of headed stud connectors. The total depth of the cross-beams is 1/15 to 1/10 of the transverse
spacing of the main girders. Shear connections (web only) are often specied for the eld splices
with the vertical stiffeners of the main girders because of the complexity and cost of exural
connections (web and anges).

When evaluating the cost of a composite deck working in single composite action, it may be useful
to have an idea of the weight qs of steelwork related to the total surface of the concrete slab. For a
twin I-girder framing system with vertical girders, the following statistical relationship may be used
in kg/m2 (Bernard-Gely and Calgaro, 1994):
qs = 0.105L1.6 + 100 (4.3)

where L (in metres) is the main span of a continuous beam with optimum span distribution.

4.1.4 U-girders for composite box girders


As an alternative to a multi-girder or a twin I-girder deck, the deck may include an incrementally
launched steel U-girder completed with a concrete slab on launch completion, to obtain a
composite single-cell box girder (Figure 4.7).

The U-girder is particularly effective at resisting the launch stresses and the weight of a forming
carriage for in-place segmental casting of the concrete slab because of the wide bottom ange and
enhanced lateral stability. The low position of the neutral axis and the wide compression ange
improve the negative bending capacity during launch, and the U-girders are therefore particularly
suitable for long-span launching with short noses (Figure 4.8).

The hollow composite section provides high torsional strength and stiffness, which are useful in
curved bridges, axial piers at the deck centreline, and slender superstructures. The resistance to
distortion is also higher and more uniform than in a twin I-girder open frame; however, torsional
and distortional strength are achieved after the onset of composite action, and temporary lateral
bracing is necessary between the top anges during the launch and construction of the concrete
slab. A composite box girder also offers good corrosion resistance (the external faces are smooth,
and the internal ones are protected) and a pleasant aesthetic aspect.

Figure 4.7 Composite box girder

235

Downloaded by [] on [25/09/17]. Copyright ICE Publishing, all rights reserved.


Bridge Launching

Figure 4.8 Incremental launching of a steel U-girder

A composite box girder is more expensive than a twin I-girder deck due to the higher quantity of
steelwork, complex workshop and eld activities, and higher shipping costs. The cost difference
can be reduced by fabricating full-width U-girders in the workshop, to avoid longitudinal eld
splices in the middle of the box. Exceptional road vehicles can reach 2527 m in length, 6 m in
width and 7080 tonnes (metric tons) in weight, and a 6 m-wide U-girder may be sufcient for
a 12 m-wide deck slab.

Unlike I-girders, U-girders have a wide and thin bottom ange, which must be stiffened to prevent
out-of-plane buckling under longitudinal compression in the negative bending regions of the con-
tinuous beam. Different congurations of the bottom ange lead to different web responses, and
tension eld theories developed for I-girders may not be readily extrapolated to box girders. The
stability of thin stiffened plates also depends on geometry tolerances, and classical elastic theory
may be inadequate in the presence of initial geometry imperfections and locked-in residual
stresses.

The thickness of the bottom ange varies longitudinally from 12 mm at the midspan to 3050 mm
in the negative bending regions. The central strip may be thinner (Figure 4.7), although this
requires thicker plates under the webs and two additional full-length longitudinal welds.
Savings in weight, reduced shear-lag effects and enhanced ange behaviour balance the higher cost
of welding.

The webs may be inclined, to reduce the width of the bottom ange and the number of longitudi-
nal stiffeners (Figure 4.9). Inclined webs increase workshop costs, especially for curved bridges,

236

Downloaded by [] on [25/09/17]. Copyright ICE Publishing, all rights reserved.


Steel and composite bridges

Figure 4.9 Internal stiffening of a U-girder with inclined webs

but improve aesthetics, reduce the shear-lag and pier-cap dimensions, and facilitate control of out-
of-plane buckling. Inclined webs also allow adjustment of the transverse spacing of the top anges
for enhanced control of distortion and optimum transverse bending in the concrete slab.

Shear lag is particularly severe in the regions adjacent to the application points of concentrated
loads, in ange plates that are wide in relation to their effective length, and in ange plates that
are longitudinally stiffened. The bridge bearings apply high localised loads, the effective length
of the bottom compression ange is about 40% of the span length, the span-to-width ratio of the
bottom ange is well within the critical range, and the plate is heavily stiffened to control out-of-
plane buckling. In a continuous box girder, therefore, the combination of these circumstances may
generate large shear-lag effects in the bottom ange regions over the supports.

237

Downloaded by [] on [25/09/17]. Copyright ICE Publishing, all rights reserved.


Bridge Launching

Figure 4.10 Concrete bottom slab for double composite action

The shear-lag effects of the longitudinal stiffeners may be avoided by casting a concrete slab onto
the bottom ange in the nal support regions of the continuous beam (Figure 4.10). The concrete
slab stabilises the bottom ange from out-of-plane buckling and avoids most of the stiffening
costs. In-place casting of the concrete slab involves low-cost eld operations, and the impact on
self-weight bending is typically acceptable. Double composite action (Rosignoli, 1998; Saul,
1996) is very effective in resisting negative bending in the support regions, and its contribution
to the exural capacity of the composite section may be extended to most of the self-weight stresses
by casting the bottom slab prior to the top slab. Open longitudinal ange stiffeners designed for
the longitudinal compressive stress of incremental launching construction are embedded into the
concrete slab on launch completion, and also work as shear connectors.

The webs of U-girders are more heavily stiffened than the webs of I-girders for the control of
deformations and distortion during fabrication and site assembly. The plate thickness adheres
to the shear demand more closely, and may be pre-sized (in millimetres) using (Bernard-Gely and
Calgaro, 1994)

BL
tw = 10 + (4.4)
100

where B is the total width of the deck and L is the span length, both in metres. The dimensions of
the top anges depend on positive and negative bending during the launch and staged placement
of the concrete slab, and on the type of connection with the concrete slab. The width of the ange
plates may vary from 600 mm for short spans to 1500 mm for long spans and wide deck slabs. The
plate thickness may reach 150200 mm in exceptional cases.

The in-plane exural stiffness of the U-section provides a degree of lateral restraint to the top
compression ange. In-plane bending is also more uniform throughout the length of the bridge
than in an open-grid deck. In spite of this, the stability of the top compression ange is typically
controlled with lateral bracing during incremental launching of the U-girder and staged construc-
tion of the concrete slab. Lateral bracing between the top anges is designed for hollow section
behaviour prior to the onset of composite action, and is typically specied also for rectilinear
bridges. For curved bridges, AASHTO (2003) species a minimum cross-sectional area of the
diagonals of lateral bracing based on the ange width. AASHTO (2003) also species the stiffness
of the lateral braces based on the torsional moment applied to the cross-section, the equivalent
plate thickness, the constant of torsion of the hollow section, and the angle between the top ange
and the lateral diagonal.

Cross-frames and diaphragms are used to minimise distortion and warping of the cross-section.
For curved bridges, AASHTO (2003) species the maximum spacing of cross-frames and

238

Downloaded by [] on [25/09/17]. Copyright ICE Publishing, all rights reserved.


Steel and composite bridges

diaphragms based on the length of the span and the radius of plan curvature. The minimum area
of the cross-frame members is determined based on the width and depth of the girder, the
longitudinal spacing of the cross-frames and the thickness of the web plates.

The concrete slab is typically supported only on the top anges of the U-girder, provided that
they are not excessively spaced. Floorbeams are used rarely because they complicate the recovery
of the central form table for in-place casting of the concrete slab. Left-in-place concrete panels for
two-layer slabs and ribbed-sheet panels are used when the inner cell of the box girder is too narrow
to operate a sliding form table. Floorbeams at 45 m spacing have been used successfully to
support full-depth precast deck panels.

The total depth of the U-girder can be pre-sized using (Bernard-Gely and Calgaro, 1994)

L 0.7
H= B (4.5)
205

where B is the width of the concrete slab and L is the length of the design span, both in metres.
Some design standards specify a maximum span-to-total-depth (steel girder plus concrete slab)
ratio of 25, and a maximum span-to-steel-girder-depth ratio of 30. The unit weight qs of steelwork
referred to the total surface of the concrete slab may be estimated in kg/m2 using the following
relationship (Bernard-Gely and Calgaro, 1994):

qs = 2.85L + 45 (4.6)

where L (in metres) is the main span of a continuous beam with optimum span distribution.

If the concrete slab is very wide, the use of twin box girders decreases the quantity of steel in the
bottom ange, avoids longitudinal stiffeners, and facilitates delivery and incremental launching of
long girder segments. This solution is often used for short spans and slender decks. The box
girders are delivered inclusive of top and bottom anges and are connected with cross-beams
on launch completion (Figure 4.11). The top anges support the forming carriage for in-place
casting of the concrete slab and avoid forming the box cells, and the cross-beams support the form
table for the central strip of the concrete slab during segment casting and repositioning of the form
table (Rosignoli, 2013). The exural and torsional stiffness of the box girders simplies launching
and does not require additional bracing.

4.1.5 Girders with an orthotropic deck


The decking system of a steel girder has multiple functions. The transfer of the vertical load to
the main girders is one of the primary functions, and providing lateral bracing and stiffness to the
main girders are also primary functions. The orthotropic deck of a steel bridge also works com-
positely with the main girders and oorbeams, and provides exural stiffness and strength.

Steel girders with an orthotropic deck are used for incremental launching construction on the
longest spans. The orthotropic deck is very light, typically of the order of 2.0 kN/m2, which offers
several advantages in bridge launching applications.

g The decking system is much lighter than a concrete slab, which results in weight saving in
the main girders. Since the orthotropic deck is also the top ange of the main girders and
oorbeams, the weight saving is reinforced (Bernard-Gely and Calgaro, 1994).

239

Downloaded by [] on [25/09/17]. Copyright ICE Publishing, all rights reserved.


Bridge Launching

Figure 4.11 In-place casting of 25 m slab segments over twin box girders. (Photo: Doka)

g Site processing of uid concrete is minimised. This is an advantage in cold-weather


applications, as steel deck panels are fabricated in a workshop, and many panels may be
stored on site by multiple stacking (Zhuravov et al., 1996). Assembly of the deck panels is
also particularly fast.
g The steel girder can be launched with the decking system in place, and only nishing and
paving are necessary on launch completion. Asphalt paving, however, poses some
challenges, as in-plane thermal gradients due to hot asphalt may generate substantial
bending.
g Modules of an orthotropic deck can be applied to the front span on launch completion, to
use the braced steel girders as a launch nose.
g The deck increases the moment of inertia for longitudinal bending during the launch, which
compensates for the extra bending due to the heavier cross-section.
g The deck increases torsional stiffness and the lateral moment of inertia, which facilitates
control of lateral exuretorsion buckling and wind-induced vibrations in the long front
cantilever.

An orthotropic deck typically includes a continuous 1214 mm deck plate that is stiffened in the
two orthogonal directions; the plate thickness can reach 20 mm in sections subject to high exural
stresses. Vertical plates or L-shapes spaced at about 300 mm may be used as open longitudinal
stiffeners; most modern bridges use 68 mm, 700750 mm-wide folded plates for closed longi-
tudinal stiffeners supporting the deck plate at 300 mm spacing. Several design standards specify
a minimum thickness and dimensions for deck plate and stiffeners. The orthotropic deck does
not participate in vertical shear transfer, and increases the weight of the steel girder during

240

Downloaded by [] on [25/09/17]. Copyright ICE Publishing, all rights reserved.


Steel and composite bridges

launching, and thicker webs may therefore be necessary for control of local buckling during the
launch.

Steel decking systems are mostly used on long spans because of high fabrication and maintenance
costs. Long-span bridges often have a streamlined cross-section with high torsional stiffness, and
orthotropic decks are therefore mostly used for box girders. Several orthotropic-deck bridges have
also been designed with twin I-girders. Three twin I-girder bridges on the Ankara Peripheral
Motorway in Turkey were incrementally launched on 147 m main spans from the opposite
abutments complete with the orthotropic deck (Popov and Seliverstov, 1998). Orthotropic decks
are also used on long-span trusses, while medium-span trusses typically use concrete slabs. Rapid
construction may justify the use of orthotropic decks on short-span trusses erected by incremental
launching (Contin et al., 2011).

Steel box girders with an orthotropic deck offer high exural capacity and torsional stiffness
during launching. When the deck of a cable-stayed bridge carries bidirectional trafc, a single
central plane of stay cables may be used, and the box girder may be launched over temporary piers
and suspended from the towers on launch completion. The webs are designed for the low shear
stresses of the nal cable-stayed span, and launching may become challenging. When a steel box
girder is used for the main span of a cable-stayed bridge and a PC box girder is used for the
approaches, a front segment of the box girder can be attached to the PC girder to be used as a
launch nose (Lockmann and Marzahn, 2009). This construction method was used for a cable-
stayed bridge over the Rhine River in Germany (Figure 2.16).

4.1.6 Trusses
Steel trusses are used on spans longer than those achievable with I-girders of reasonable depth.
Short-span trusses are rarely used for bridge structures because steel trusses require more hand
welding and involve higher fabrication and erection costs than braced I-girders. Modern castel-
lated trusses offer lower fabrication costs, simpler welding with prole-tracking equipment and
no gusset plates at the eld splices; in spite of that, however, they are still more expensive than
braced I-girders.

A steel truss may have a steel orthotropic deck or a concrete slab. Several varying-depth trusses for
highway and railway bridges have been erected as balanced cantilevers. Negative bending charac-
terises this construction method, and some trusses therefore have a bottom concrete slab in the
pier regions for double composite action. The use of a top concrete slab is the typical solution for
medium-span trusses because of cost reasons.

A very small number of steel trusses have been launched complete with a steel orthotropic deck.
The 814 m-long Po River Bridge in Italy includes 11 spans, each 75 m in length. The W-shaped
cross-section includes two tubular V-trusses braced at the bottom and supporting a 14.5 m-wide
steel orthotropic deck with omega-stiffeners (Contin et al., 2011). The truss was launched over the
historical piers of a previous truss bridge destroyed during World War II.

The 2600 m-long Parana River Bridge in Brazil includes two Warren trusses with verticals. The
trusses carry a 17.4 m-wide roadway platform on the concrete slab at the top chord level, and a
single-track railway supported on the bottom chords between the trusses. Two sections of the
truss, 600 and 700 m in length, respectively, were incrementally launched from the opposite river-
banks on 100 m spans without the use of a launch nose (Malite et al., 2000).

241

Downloaded by [] on [25/09/17]. Copyright ICE Publishing, all rights reserved.


Bridge Launching

Figure 4.12 Full-span launch of a Vierendeel truss on 65 m spans. (Photo: Knight)

A built-up weathering steel Vierendeel truss launched over 11 electried railway tracks and the
platforms of the Stratford Regional Station in London, UK, shows the potential of incremental
launching construction in congested urban areas (Figure 4.12). The open truss without diagonals
is well suited to the full-height glazed facade that provides edge protection against falling objects.
The 130 m curved truss was assembled on an aerial platform and launched over two 65 m spans.
The deck is 12 m wide and the depth of the edge trusses ranges from 4.5 m at the bridge ends to
6.5 m at the midspan. The inner truss is vertical at the bridge ends and progressively warps to a 158
outward inclination at the centre of the bridge (Knight and Fryer, 2011). A trussed nose and a rear
tail frame for truss stability were used for full-span launching.

4.2. Girder segmentation


Incremental launching construction of a steel girder and a PC girder involve very different activi-
ties. A PC girder is cast in-place from loose materials (uid concrete, reinforcement, tendon
anchorages, ducts, strand and scuppers). Even when the deck is made of precast segments, these
are built from loose materials without major dimensional restraints. Handling and shipping of
precast segments are subject to weight and dimensional restrictions, but standard 3 m segments
are generally easy to deliver.

The design of a steel girder is inuenced by fabrication requirements from the very beginning. The
steel plates are rolled in a foundry, transported to a workshop, cut into pieces and welded to form
subassemblies, transported to a paint shop, painted, shipped to the eld, assembled and launched
into position. Each of these steps involves dimensional restraints, which explains the importance of
assembly techniques and their inuence on the total cost of the bridge.

242

Downloaded by [] on [25/09/17]. Copyright ICE Publishing, all rights reserved.


Steel and composite bridges

Standard dimensions of the steel plates vary from country to country. The maximum plate width is
determined by the roll width in the rolling mill, and is generally about 4.5 m in Europe. The
maximum plate length and thickness depend on the volume of the ingot to be rolled, which in turn
depends on the foundry. The maximum commercial length can exceed 30 m for plates thinner than
20 mm and 1.53.8 m wide. The maximum thickness is not limited by rolling operations but
depends mainly on welding requirements: 150 mm is generally an upper threshold. The maximum
plate weight is 1520 tonnes, and the minimum weight is 35 tonnes.

Steel plates are shipped via truck or rail wagon. The maximum standard dimensions are about
24 m in length and 2.7 m in width, which sometimes increase up to 33 m and 2.8 m, respectively.
Commercial dimensions are signicantly inuenced by the standards of individual countries.

Steel plates are worked to create bridge segments, the features of which depend on their handling
in the workshop. Workshop organisation has a signicant inuence on the fabrication cost of the
bridge and affects its design. The fabrication sequence is generally as follows (geometry and weld
controls are distributed along the whole fabrication process):

1 Fabrication preparation: programming of CNC equipment, planning of assembly


sequences, organisation of production and quality control.
2 Working of the individual components: piece marking, scribing of the piece prole and
positional references, cutting, grinding and edge preparation, drilling of holes, sandblasting
and cleaning.
3 Assembly of segments: positioning of the individual components, locking and tack welding
of plates, edge pre-heating, welding, straightening after welding.
4 Trial assembly, which is often combined with the drilling of bolted eld splices: a few
small-diameter bolts are drilled to block the splice, alignment pins are used to force the
segments into position with the desired angle break, the holes for the bolts are CNC drilled
with the nal splice geometry, the holes for the small temporary bolts are widened to the
nal diameter to complete splice preparation.
5 Segment completion: welding of stiffeners, minor plates and headed stud connectors,
machining of surfaces at butted joints.
6 Final cleaning and painting.

The segments are shipped to the erection site via a roadway or railway. River or sea transportation
of macro-segments is adopted when possible. Many closure spans of portal frames and long-span
bridges have been assembled off site, barged under the bridge and strand jacked into position.
Truck delivery is often the most exible solution, as it allows most sites to be reached; the most
practical weight of the segments is 2030 tonnes, depending on the type of trailer and the form
of the segment. The maximum length is about 22 m; the maximum width is about 3.20 m if loaded
horizontally and 4.50 m if loaded vertically.

Exceptional means of transport can exceed these weights and dimensions. When the length is lead-
ing, the segment can reach 3335 m in length, 3 m in height, 3 m in width and 7080 tonnes in
weight. When the width is leading, the segment can reach 2527 m in length, 56 m in width and
7080 tonnes in weight. Railway shipping is subject to more stringent restrictions, as the segments
must be placed on standard wagons: the maximum length is about 32 m, the height is about 3 m,
the width is about 2.5 m and the weight is about 50 tonnes. Interest in railway shipping also
depends on the possibility of unloading wagons at a station close to the erection site.

243

Downloaded by [] on [25/09/17]. Copyright ICE Publishing, all rights reserved.


Bridge Launching

The design of the eld splices also affects deck segmentation. Slip-critical bolted connections are
preferred when simplifying and accelerating eld assembly are prime concerns. These joints
weaken the cross-section, and are therefore placed as close as possible to self-weight counter-
exure points. Full-penetration butt weld joints are preferred for aesthetic quality, no maintenance
and minimal interference with the launch bearings. Since welding does not affect the exural
capacity of the cross-section, welded splices are distributed freely within the structure.

Welded eld splices are normally located on a single plane, as staggering the welds of anges and
the web does not improve the performance of the connection. Preparing and preheating the joints
and maintaining proper alignment is also easier when all the welds are in the same plane. Fillet
welds of anges to the web may also be extended to the very end of the girder, which simplies
handling and provides better stability when the anges are locked for welding.

From an organisational point of view, eld assembly should be simplied as much as possible.
Segment alignment, joint welding and quality control of the connections are more difcult and
expensive on site than in a workshop. This suggests subdividing the steel girder into segments
as long and heavy as possible, compatible with the handling capability of the workshop and
assembly yard, the available transportation means and the camber design.

Considerable analytical work has been done in the area of optimisation of plate girder design
(Xanthakos, 1994). The typical approach is based on relating the cross-sectional properties to a
given set of bending moments and shear forces. The web depth is assumed constant, and the
cut-off points in the anges and the minimum length of their plates are decided from practical con-
siderations. Each span is divided into girder segments based on the number of plate cut-off points
and the bending diagram. For a span composed of n girder segments, the bending distribution is
known once the maximum moment Mi has been determined for each segment.

Either the cost or the weight of the girder may be considered as the objective for the optimisation
process. Most investigators choose the weight, but the same approach can be followed on a cost
basis. Starting from a symmetrical I-girder with a constant geometrical depth h of the webs
(distance between the ange plate centroids) and a thickness tw but with varying ange plates, the
total weight Q for a span L is
 n 
Q  g htw L + 2Af,i bi (4.7)
i

where g is the specic weight of steel, Af,i is the cross-sectional area of a ange segment and bi is the
length of the ange segment. If the design of the girder is optimised for bending, the section
modulus Sx may be related to the SLS axial stress limit s specied by the design standards. A
similar approach may be followed for the ULS exural capacity based on the slenderness class
of the cross-section:

tw h2 Mi
Sx,i  Af,i h + = (4.8)
6 s
Extracting Af,i and combining gives
   
n  
 2 Mi bi  htw bi
Q  g htw L + (4.9)
i
hs 3

244

Downloaded by [] on [25/09/17]. Copyright ICE Publishing, all rights reserved.


Steel and composite bridges


The weight of a girder with a chosen web thickness is a minimum when the value |Mi bi| is a
minimum. If this criterion is applied to all the spans of the continuous beam, the summation

|Mi bi|j of the values of |Mi bi| for the j spans must be a minimum to yield the minimum
girder weight.

A two-stage iterative procedure is used to minimise |Mi bi|j . The rst step is to determine the
optimum location of the cut-off points based on the number chosen and the camber requirements
related to the elevation and vertical deection. Angle breaks are typically used at the eld splices,
and the anges are kept straight in between. The resulting ange plate lengths bi are independent of
the optimum web height, which is determined later. In the rst stage of the analysis, the design
values of bending Mi are determined for the assumed lengths bi . In the second stage, the moments

are kept unchanged and the segment lengths are varied to minimise |Mi bi|j . The new values
of bi are fed back into the rst stage to recalculate the moments, and the procedure is repeated until
Mi is compatible with bi .

The choice of the web may be based on the same procedure. Using a modied section modulus
Sx,m for the continuous beam as a whole instead of the section modulus required for the maximum
moment,
 
Mi bi j
Sx,m = (4.10)
s Lj

where Lj is the total length of the j spans,
then setting dQ/dh = 0 allows the web depth for a

chosen thickness tw to be obtained as h = 3Sx,m /tw . The depth-to-thickness ratio is nally
checked against the criteria stipulated by the design standards. This optimisation method for
symmetrical I-girders can be easily adapted to asymmetrical I-girders with different ange plates,
to U-girders to be completed with a steel orthotropic deck or a concrete slab, and to hybrid girders
using conventional-grade steel in the webs and high-grade steel in the top and bottom anges in
the negative bending regions and in the bottom anges in the positive bending regions.

Regarding the bending moment to be used for the calculations, a steel girder built by incremental
launching is affected by three states of stress.

g The launch stresses depend on the self-weight of the steel girder, the localised load of the
pier diaphragms, the movement of a non-launchable cambered prole over aligned launch
bearings, the effects of geometry imperfections and misaligned launch bearings, and
meteorological loads such as thermal gradients and wind. After enveloping the self-weight
launch bending with aligned bearings, the effects of camber, geometry imperfections,
bearing misalignment and meteorological loads are calculated in some characteristic
sections and superimposed.
g The slab construction stresses depend on the segmental erection procedure. When the
concrete slab is cast in-place, the construction stresses are calculated by applying the weight
of the concrete slab and forming carriage to the steel girder according to the segmental
casting sequence. The progressive onset of composite action may be disregarded when
optimising the girder segmentation. The weight of special construction equipment is
minimal when the slab is incrementally launched over the steel girder, assembled with full-
depth precast panels erected with ground cranes, or cast in-place over precast deck panels.
When the steel girder supports an orthotropic steel deck, the deck is applied prior to
launching, and no slab construction stresses are considered.

245

Downloaded by [] on [25/09/17]. Copyright ICE Publishing, all rights reserved.


Bridge Launching

g The service stresses depend on the self-weight of the composite girder, the superimposed
dead load, thermal stresses, shrinkage stresses and live loading. Composite action is
considered in the positive bending regions of the continuous beam, and the longitudinal
slab reinforcement is considered in the section moduli of the negative bending regions.

The three bending moment diagrams in the steel girder are superimposed, and girder segmentation
is optimised for their envelope. When the nal distribution of the cut-off points has been deter-
mined, the length of the segments to be delivered to the assembly yard is chosen based on techno-
logical requirements and the lifting capacity of the handling equipment. The U-girders for steel
and composite box girders may be transported full-width when not excessively wide; otherwise,
two L-girders may be welded longitudinally or may be completed with a central bottom ange
panel welded longitudinally to the L-girders. The orthotropic panels for the top deck and the bot-
tom ange of box girders have different modularity. The panels are typically submultiples of the
web modules, and are spliced by eld welding of ange plates and stiffeners.

4.3. Organisation of the assembly yard


When the bridge is short, the girder is assembled behind the abutment and launched into position
in one operation. This is the typical approach for single spans and tied arches. The steelwork is
often launched with winches, reeved ropes and safety brakes, and the concrete slab is cast in-place
on launch completion. Two xed front bogies are used at the launch abutment, and two rear
mobile bogies rolling on runway beams support the rear end of the span during launching. The
span may be ballasted for control of overturning or may be launched over temporary piers.
When the span is ballasted and can partially overhang beyond the launch abutment, the launch
nose may be hoisted from the opposite abutment and connected to the span for launch completion
without temporary piers (Romaro and Romaro, 2000).

When the bridge is longer, the launch is performed in a series of increments, and the eld activities
are divided into two sets of operations alternated in time: the assembly of a new section of girder
behind the completed girder, and the launch of the whole girder over the piers.

The assembly area is often twice as long as the end span of the bridge, to control overturning prior
to nose landing at the rst pier. When site restrictions limit the space available, the rear end of the
girder is ballasted to control overturning, or a temporary pier is used in the end span of the bridge.

The assembly area may be designed to contain three girder segments. The segments are as long as
possible to reduce the number of eld splices, and their length is kept constant to avoid multiple
foundations for the assembly supports and moving them back and forth within the assembly area.
During positioning and geometry control of the rear segment, the central segment is spliced or
welded, and the front segment is nished and painted at the eld splices. Three specialised working
areas (assembly, splicing and nishing) are set up at xed locations, the launch steps are as long as
the segments, and the duration of the activities on the three segments governs the time schedule
of assembly. The launches may also be as long as two or three segments; in this case, the eld
activities interfere with each other and must all be completed prior to launch, but disruption of
the assembly due to launch operations is lessened.

Handling long and heavy girder segments requires special equipment. When the girder is
assembled within the project right-of-way, the segments are often stored in a buffer area. The need
for easy access and manoeuvring of delivery trucks moves the storage area far from the assembly

246

Downloaded by [] on [25/09/17]. Copyright ICE Publishing, all rights reserved.


Steel and composite bridges

area, the need for storing several segments requires easy access to the storage platforms, and the
distance between the storage and assembly areas requires fast and powerful lifters able to shuttle
back and forth between the two areas.

Crawler cranes can be used to unload the segments from the delivery trucks and to crawl with the
segments to the assembly area. This method is not the most efcient, but is sometimes used when
the layout of the storage and assembly areas can accommodate large runways for travel and the
swing radii of the crane. Crawler cranes are very slow for these operations (Rosignoli, 2013).

If the assembly yard is narrow and rectangular, one or two rail-mounted portal cranes are often
the best solution. The portal cranes require narrow runways along the rear storage area and the
front assembly area. The girder segments are stacked for optimal storage efciency and are lifted
over stored segments during delivery to the assembly area. More than 300 tonnes of steel have been
erected per month with this type of equipment (Zhuravov et al., 1996). Rail-mounted cranes
require full-length runway beams and at terrain. The power supply is typically through retract-
able cables, but long runways may suggest the use of carried power-pack units. Three-hinge portal
cranes typically lift the segments with hydraulic winches. Portal cranes are less expensive and
easier to operate than the straddle carriers but do not offer the same manoeuvring capability.

If the storage area is square or irregular, a motorised straddle carrier is often preferred. The
straddle carriers are off-road lifting vehicles that are used to lift girder segments from the delivery
trucks, to position them onto the storage platforms, to pick them up for nal delivery, to move
them to the assembly area and to place them onto the assembly supports (Figure 4.13). These

Figure 4.13 Straddle carrier for the handling of segments

247

Downloaded by [] on [25/09/17]. Copyright ICE Publishing, all rights reserved.


Bridge Launching

wheeled cranes provide unrivalled exibility of operation but require large runways. Custom lifting
beams are used to pick up the segments with the desired suspension conguration. The straddle
carriers are not very tall, as the girder segments are assembled proceeding backward from the rear
end of the completed girder.

The straddle carriers are capable of relatively high speeds, up to 30 km/h, and are available in a
variety of congurations with a load capacity ranging from 50 to 150 tonnes. Good manoeuvra-
bility and the ability to operate on typical heavy-vehicle roads provide exibility to the organisation
of the assembly yard. Operation with multiple stockyards with linear or herringbone layouts and
complex routings is possible, giving straddle carrier-based transport systems unrivalled exibility
(Rosignoli, 2013).

The straddle carriers handle girder segments quickly and effectively. Rapid loading, transporting
and unloading and the ability to self-load and unload minimise delays and boosts carrier utilisation.
Hydraulic winches or long-stroke double-acting cylinders are used to lift the load. Integrated
position sensors provide cylinder position feedback for closed-loop automatic synchronisation
of multiple hoist cylinders. Variable displacement hoist pumps supply pressure and ow to operate
the hoist and swing cylinders. Counterbalance valves are mounted internal to the hoist cylinders
for load-holding capability.

An independently controlled steering cylinder is mounted at each wheel. Redundant internal


position transducers are installed in each cylinder to provide feedback to the steering computer.
The straddle carriers typically offer four steering modes: four-wheel coordinated steering, two-
wheel (front) automotive steering, two-wheel (rear) forklift steering and four-wheel parallel (crab)
steering. The steering computer synchronises the wheels during manoeuvring and facilitates access
to storage platforms and assembly supports. Four-wheel steering and four-wheel drive are helpful
in negotiating tight turns. Soil preparation, well-maintained runways and accurate drainage are
necessary to avoid premature wear and tear of the steering and drive mechanisms of the crane.

When the bridge is short, the girder is entirely assembled behind the abutment and launched in one
operation. Savings result from fewer alignment and launch operations and lower labour demand.
Many assembly supports are expensive, and they are therefore designed with modularity to reuse
them on new projects or to resell them at the end of the launch. Additional costs result from the
earthworks for a long assembly yard, especially when the girder is launched at its nal elevation
and the assembly yard is therefore at a lower level than the bridge bearings. Launching the girder
from the nal elevation of the access embankment accelerates bridge completion and facilitates
rainfall drainage, but the launch bearings at the piers must be supported on temporary towers
as tall as the abutment back-wall. The temporary towers are designed for stepped lowering of the
girder on launch completion.

The typical assembly yard is as wide as the steel girder plus a 34 m lane on either side for access
and the operation of cranes or straddle carriers. The launch surface of the steel girder is typically at
a higher elevation than the permanent bearings, as the launch bearings are taller than the latter
and supported on the same pedestals. The exibility of the steel girder facilitates jacking, removal
of the launch bearings, and stepped lowering onto the permanent bearings on launch completion.

Each segment of the girder is positioned onto adjustable assembly supports located at the quarters
of the segment length, to control deections and rotations of the end sections, and to facilitate

248

Downloaded by [] on [25/09/17]. Copyright ICE Publishing, all rights reserved.


Steel and composite bridges

Figure 4.14 Adjustable assembly supports for L-girders with a central bottom plate

access to the eld splices in the bottom anges. The assembly supports of the I-girders are aligned
under the webs. Two additional support lines are necessary when a U-girder includes two outer L-
segments and a central segment of the bottom ange, and the assembly supports are widened to
also support the central segment (Figure 4.14). The L-segments are positioned rst, and then the
bottom ange panel is positioned and welded.

The assembly supports are set to the design camber for elevation and vertical deection; lateral
alignment occurs contemporaneously in curved bridges. Then, the girder segments are positioned
onto the supports and adjusted to the butt weld clearance for the eld splices, or to align the holes
of bolted connections. This recreates the geometry conditions of the workshop trial assembly
and results in accurate alignment of the pier diaphragms at the end of the launch. After minor
geometry corrections by jacking the segments or by pulling bars anchored to the segments,
hydraulic clamps are applied to lock the edges of the segments at the eld splices in preparation
for welding. Minor adjustments of the cross-sectional geometry are still possible in this phase.

The geometry tolerances allowed by the design standards may be excessive for the bottom anges
of launched bridges, as defects in planarity in the sliding surfaces may complicate launching and
cause stress concentrations. BS 5400 (BSI, 1982) accepts unintended deviations from planarity due
to plate misalignment in a welded butt joint when these do not exceed the lesser of 0.15 times the
plate thickness or 3 mm. A sharp 3 mm step in the launch surface may cause many problems when
passing over sliding launch bearings. Irregularities in the sliding surfaces must therefore be
detected, attened and repainted prior to launching.

Bolted eld splices are often avoided in the main girders because of design restrictions on the
optimum girder segmentation and the interference of full-width lap plates with the launch

249

Downloaded by [] on [25/09/17]. Copyright ICE Publishing, all rights reserved.


Bridge Launching

Figure 4.15 Welding sequence for field splices

Flange

Web

bearings. With progress in welding technology, eld splices in the main girders are welded increas-
ingly frequently. A welding sequence for the eld splices (Bernard-Gely and Calgaro, 1994) is
shown in Figure 4.15. It is based on alternately welding anges and web for a portion of their
thickness.

1 Weld the central portion of the double V-joint in both anges (one-third to one-half of the
plate thickness) full width.
2 Weld a portion of the thickness of the single V-joint in the web (about one-half of the plate
thickness) full depth.
3 Complete welding of the anges.
4 Complete welding of the web.

For deep webs, vertical welding may be divided into two or more sections, to ensure a more
uniform transverse shrinkage of the joint. These staggered sequences generally ensure adequate
control of thermal shrinkage of welds.

Most butt joints in the eld splices of the webs are the single V-type. For thick webs (above
14 mm), a double V-joint may be specied to reduce the amount of welding and the angular
distortion due to thermal shrinkage. The design of the ange joints to be eld welded depends
on the plate thickness. A double V-joint with half of the welding on either side inhibits thermal
distortion but requires considerable overhead welding. For this reason, an asymmetrical double
V-joint may be specied so that 6070% of the ange thickness is welded from the top, and the
remainder from the bottom.

Automatic welding with prole-tracking equipment may be used for the longitudinal eld splices
of U-girders that include two symmetrical L-segments. In wide bridges, shipping restrictions may

250

Downloaded by [] on [25/09/17]. Copyright ICE Publishing, all rights reserved.


Steel and composite bridges

require including a central bottom ange plate between the two L-segments, and two full-length
longitudinal eld welds are necessary. Robotised welding requires additional eld assembly
operations for pier diaphragms and transverse stiffeners due to the clearance requirements of the
welding unit. Transverse eld splices are typically hand welded.

The eld welds are cleaned, attened and controlled, and paint is applied to the eld splices so that
the steel girder is launched without any need for additional work. Temporary bracing to be
removed on launch completion or after casting the concrete slab is bolted. Slip-critical bolted
connections are also used for eld splices subject to fatigue and for minor facilities such as inspection
catwalks and supports of utilities or drainage pipes.

A congested urban environment or an insufcient area behind the abutment may require the
fabrication of girder subassemblies on the ground and crane lifting or strand jacking into position
for launching. The segments of twin I-girders for the approach railway viaduct of the Tagus
Suspension Bridge in Portugal (Reis, 2001) were shipped to the site in half-sections to be welded
at the diaphragms. The 919 m girder was launched from the opposite ends; a 392 m section was
rectilinear, a 527 m section had a constant 1000 m plan radius and an expansion joint was kept
between the two sections. Launch platforms 76 m in length supported on the piers and propped
from the foundations were used to support the segments at the launch elevation during connection
to the rear end of the launched girder. A similar organisation of the assembly yard was used for the
Dunaujvaros Danube Bridge in Hungary (Horvath et al., 2006).

The complexity of site organisation increases with the number of activities to be performed.
Incremental launching of the 814 m three-dimensional space frame of the Po River Bridge in
Italy required moving 11 spans of space frame between different fabrication areas, lifting the spans
to the launch elevation and skidding them to the rear end of the launched bridge for welding. Five
working areas were set up for span fabrication. Two assembly lines were used to fabricate 15 m
panels of trussed webs, a line was used to assemble 15 m bridge segments, a line was used to weld
the 15 m segments into 75 m spans and the fth line was used for welding the new span to the rear
end of the launched bridge. The area between the nal welding facility and the launch abutment
was used for quality control and metallisation and painting of the space frame (Contin et al.,
2011). Only ange modules and diagonals were prefabricated off-site. More than 100 welders and
ve quality control teams worked on site during the bridge construction.

On assembly completion of the new group of segments, launch bearings installed on hydraulic
jacks are lifted to support the girder, and the assembly supports are lowered in preparation for
launch. Numerous solutions are available to support the girder in the assembly yard during
the launch. Launch plates and Neoon pads are rarely used because of the need for elevation
adjustment in relation to the camber of the girder. Mobile supports used for assembly and launch,
hydraulic support saddles (Hoeckman, 2001) and combinations of these systems have been used
for several launches, depending on experience, availability of equipment, and the eld splice design
for the bottom ange.

Cambers and deections due to positive bending in the end span of the bridge often cause girder
uplift at the launch bearings in the assembly yard. This is rarely an issue for the girder because of
the high capacity-to-demand ratio on such short spans, but the rollers and foundations of the
launch bearings should be designed considering the redistribution of support reactions due to
uplift.

251

Downloaded by [] on [25/09/17]. Copyright ICE Publishing, all rights reserved.


Bridge Launching

4.4. Launch of steel girders


The steel girder of a composite bridge is designed to carry the self-weight and the weight of the
concrete slab and casting equipment prior to the onset of composite action. Superimposed dead
load, live loads and the time-dependent effects of concrete creep and shrinkage are carried with
composite action. Although the launch stress distribution is different from the service stress
distribution, the girder is launched without the concrete slab, and this creates ample margins of
strength. In most cases, therefore, incremental launching construction requires only minor
adjustments to a steel girder designed for the nal service conditions.

Assembling and launching a rectilinear grillage of I-girders involve simple and intuitive
operations. Launching a curved U-girder does not involve particularly complex operations either,
provided that the radius of plan curvature is constant and the deck cross-fall is attained with a deeper
outer web rather than by rotating the bottom ange in the transverse plane. Segment geometry is
repetitive, and this also facilitates the analysis of the launch stresses.

In most cases, a curved girder is fabricated from straight segments jointed to shape. The segments are
designed to be as long as possible to minimise eld splicing, and the lateral offset of the webs from the
circular launch trajectory can be signicant. Keeping the support reactions aligned with the webs
requires particular care during the launch, especially when the plan radius is shorter than 500 m.

The launch nose is generally short or very short, and the stability of the long front cantilever from
lateral torsionexure buckling often requires lateral braces at both anges and closely spaced
cross-frames. Additional vertical stiffeners are often necessary to control local buckling in the web
region over the lead pier at the nose landing at the next pier, especially when a front cable-stayed
system is used for launching. Misaligned launch bearings and geometry imperfections in the steel
girder may cause local buckling in the cross-frames, which may trigger catastrophic lateral torsion
exure buckling of the front cantilever. The absence of deformations in the lateral braces and cross-
frames is therefore checked for frequently during launching, and reasons for out-of-at distortion
are investigated to avoid stress concentrations and risks of instability.

When the plan curvature varies or the rectilinear segments of a curved girder are long, the launch
bearings are placed on skidding shoes to allow lateral displacements of the support points.
Compared with a PC bridge, the support reactions are much smaller, and light cross-beams or
steel brackets are applied to the pier caps to accommodate large transverse displacements of the
launch bearings.

4.4.1 Launch bearings


A PC deck is a heavy and stiff structure. High support reactions require launch bearings with
large sliding surfaces, the small exural rotations at the supports can be absorbed with laminated
elastomeric pads and do not require articulated launch bearings, and the thickness of the
launch bearings is similar to the thickness of the permanent bearings, so that minor deck jacking
is necessary for their replacement at the end of the launch.

A steel girder is much lighter and more exible, and the webs are much thinner. The dispersal of
the support reactions within the webs requires long support surfaces, and large exural rotations
of the girder at the supports dictate the use of rocking bearings. Articulated launch bearings are
much deeper than the permanent bearings, and major jacking operations are necessary at the end
of the launch to lower the girder onto the permanent bearings.

252

Downloaded by [] on [25/09/17]. Copyright ICE Publishing, all rights reserved.


Steel and composite bridges

Further complication derives from the vertical prole of the girder. Cambers are necessary to
compensate for girder deections due to self-weight and the weight of the concrete slab, and since
the girder is exible and the concrete slab is 34 times heavier, cambering is often signicant.
Cambers are mostly attained by assembling straight girder segments to shape, and therefore the
anges have angle breaks at most eld splices. The angle breaks in the bottom anges are convex
in the nal support regions, and when they pass over the launch bearings they tend to localise the
support reaction. Combining localised support reactions, vertical shear, longitudinal axial
compression due to negative bending and the absence of vertical stiffeners may cause yielding and
local buckling in the bottom ange and the bottom web region.

For all these reasons, the launch bearings for steel girders are more complex than those for PC
bridges. A launch bearing for a PC deck is a steel plate covered with a polished stainless-steel sheet
and supported on an elastomeric bearing, and its total thickness rarely exceeds 100 mm. A launch
bearing for a steel girder is a complex mechanical device and its depth can exceed 1 m, the launch
surface is higher than the nal elevation in most cases, and the steel girder must be lowered onto
the permanent bearings on launch completion.

The launch bearings are also longer than those for PC bridges, to facilitate dispersal of the support
reactions within the webs and to control bucking of the web panels with a long patch load. The
design load of the bearings is determined by taking into account bearing misalignment due to
installation tolerances, differences in elevation in the bottom anges due to fabrication and
assembly tolerances for the steel girder, and angle breaks in the bottom anges due to cambers
and the vertical prole.

The rst type of launch bearing is based on cast-iron rolls assembled on inverted equalising beams
that balance the load in the rolls (Rosignoli, 2013). The rolls support the bottom ange of the steel
girder under the web (Figure 4.16). The rolls are 80120 mm thick, the diameter ranges from 300

Figure 4.16 Eight-roll articulated bogie negotiating a lap plate. (Photo: LaViolette)

253

Downloaded by [] on [25/09/17]. Copyright ICE Publishing, all rights reserved.


Bridge Launching

to 500 mm and the load capacity is 0.40.5 MN. The number of rolls depends on the vertical load
and their diameter, and varies between two, four and eight, with progressively higher costs. Four-
roll assemblies on three equalising beams are the typical solution for reducing the diameter of the
rolls and lengthening the patch load applied to the girders.

The bogies are articulated to cope with exural rotations in the steel girder, launch gradient, and
steps and angle breaks in the bottom ange. A roll acting against the side of the bottom ange
typically provides lateral guidance. According to CNR 10011 (CNR, 1988), the linear contact
pressure transferred by a roll with a radius r and a transverse contact breadth b under a vertical
load F is

0.18EF
slc = (4.11)
rb

CNR 10011 (CNR, 1988) limits the SLS linear contact pressure to slc 4sall , where sall is the SLS
axial stress limit, and the factored ULS linear contact pressure to gslc 4fu , where fu is the tensile
strength of steel. The linear contact pressure for a pair of rolls is slightly higher to allow for the
imperfect load distribution:

0.20EF
slc = (4.12)
2rb

In the case or four or more rolls it is



0.24EF
slc = (4.13)
nrb
where n is the number of rolls. The rolls are mounted on the equalising beam with bearing brasses
or roller bearings to reduce friction. The equivalent coefcient of friction with roller bearings is
13% in ideal conditions, but 5% is typically used for design purposes.

This type of launch bearing must be modelled accurately to analyse the dispersal of the support
reactions within the webs. Figure 4.17 shows the vertical axial stress generated by four-roll bogies
with three articulated equalising beams and by the mast of the front cable-stayed system during the
launch of a three-girder grillage. In the model, the bottom equalising beam is supported on a
spherical hinge to allow rotation of the bottom ange of the main girder about the longitudinal
axis.

The distance between the wheels is optimised by trial and error to minimise the vertical axial stress
peaks at the desired elevation in the web panel. Conceptually similar launch bearings can be
obtained by assembling two commercial rollers on articulated beams (Figure 2.38). The rollers
ensure longer patch loads than the cast-iron wheels, but the longer the roller, the less uniform the
contact pressure (Granath, 1998, 2000; Granath et al., 2000). The rollers are also less adaptable to
angle breaks and irregularities in the bottom ange.

Cable bearings are rarely used for cost reasons. In these devices, the support reaction is distributed
by a tensioned ring-cable that supports the rolls. The cable bearings are thinner and more stable
than the articulated bogies, and their robustness encourages their use in the most delicate
launches. When launching occurs along a curve, cable bearings may be placed onto skidding shoes

254

Downloaded by [] on [25/09/17]. Copyright ICE Publishing, all rights reserved.


Steel and composite bridges

Figure 4.17 Vertical load dispersal of four-roll articulated bogies with three equalising beams

for orientation and lateral alignment under the webs. Cable bearings on hydraulic jacks permit
weighing and adjustment of the support reactions and enhanced control of local buckling in the
steel girder.

Fabrication tolerances in the main girders and cross-frames can modify the support reactions so
much that one girder uplifts from the bearing, and the adjacent launch bearings resist higher support
reactions. Flange uplift is unusual with braced I-girders and accurate positioning of the launch
bearings, but less infrequent with multi-girder systems and when narrow box girders are launched
complete with an orthotropic steel deck. Overloading does not affect the launch bearings much, as
these devices are typically oversized. Load redistribution, however, modies the launch stresses in
girders and cross-frames, and requires the accurate assessment of strength and local stability.

Sliding bearings equipped with PTFE skids can be used as an alternative to roll-based bearings for
low-friction launching. Compared with a PC deck, the support reactions are smaller, launch friction
is rarely critical for the permanent piers, and the steel girder may often slide on lubricated PTFE
skids without interposition of polished or chromed stainless-steel sheets. Sliding paint may be
applied to the bottom surface of the anges to further reduce friction (Mato and Santos, 2007).
Heavier girders may require the use of Neoon pads or plywood or nylon plates with greased
anti-friction materials, which are inserted between the girder and stainless-steel launch bearings,
and are expelled in the launch direction. Launch plates of different thicknesses may be used
for compensation of the support reactions. The friction coefcient of these systems varies in the
510% range.

255

Downloaded by [] on [25/09/17]. Copyright ICE Publishing, all rights reserved.


Bridge Launching

Figure 4.18 Sliding bearing with elastomeric blocks lodged in a rocking frame. (Reproduced with
permission from ASCE)

Multi-layered elastomeric blocks covered with a dimpled PTFE plate are used to lengthen the
patch load under the web. The elastomeric blocks are 100200 mm thick: this does not increase
the total cost of the launch bearings much, and improves the transverse distribution of the support
reactions and their dispersal within the webs. The blocks are made with low-hardness thermo-
setting rubber and are lodged within rocking frames placed on ball-and-socket articulations,
transverse pins or rubber bearings (Figure 4.18). In the most delicate cases, the elastomeric blocks
may be placed onto spherical springs lodged within the rocking frame (Hoeckman, 2001).

Rotation of the rocking frame is indispensable for the high support reactions of steel girders
launched complete with an orthotropic deck, especially when a launch on temporary piers requires
minimal load eccentricity. Heavy loads may also require the reduction of sliding friction by insert-
ing polished stainless-steel sheets between the steel girder and the elastomeric blocks in the rocking
frame (Figure 4.19). Base rotation of the rocking frame may be superuous when the support
reaction does not exceed 2.0 MN per girder (Zhuravov et al., 1996).

The contact pressure on the elastomeric blocks may be higher than in the thinner Neoon pads for
PC bridges, as multiple elastomer layers facilitate the control of the stress peaks due to local
defects in the sliding surface. The vertical pressure on the elastomer governs the design of the
sliding bearings, as unconned PTFE can carry compressive stress higher than 30 MPa.
Stiffness and cost limit the length of the rocking frames, the elastomeric blocks are often wider
than the cast-iron wheels of the articulated launch bogies, and their width is considered in the
design of bolted eld splices in the bottom ange to ensure a ush central strip of adequate width.
The shear deformations of elastomer due to sliding friction are rarely an issue, as the rocking
frame supports the blocks laterally. Lateral guides equipped with low-friction metallic surfaces are
applied to one or both sides of the rocking frame to guide the bottom ange of the girder during
the launch and to resist lateral loads (Figure 4.20).

256

Downloaded by [] on [25/09/17]. Copyright ICE Publishing, all rights reserved.


Steel and composite bridges

Figure 4.19 Long rocking frame and polished stainless-steel sheets

Figure 4.20 Upper view of the sliding bearing in Figure 4.18

257

Downloaded by [] on [25/09/17]. Copyright ICE Publishing, all rights reserved.


Bridge Launching

Figure 4.21 Offset launch of a clothoid spiral. (Photo: Schimetta)

In the current state of practice, roll bearings and sliding bearings are complementary. Sliding
bearings are suited for the slow launching of heavy loads, while roll bearings are suited for medium
loads and fast launching. Both types of bearing may be placed on skid shoes for lateral displace-
ment. Skidding may be necessary for the general lateral displacement of the girder or for launching
a girder with a non-launchable plan geometry (Geier, 2010). Lateral offsets of 2.75 m from the pier
centreline were necessary for the launch of the clothoid spirals of the Auenbach Bridge in Austria
(Figure 4.21).

Both types of launch bearings are thicker than the permanent bridge bearings, and the girder must
therefore be lowered onto the bearings on launch completion. The girder is jacked, the launch
bearings are removed, the bearing pads are cast, the permanents bearings are aligned and grouted
to the pads, the girder is lowered on the bearings, the support reactions are weighted and corrected
with shims, and the pier-cap gear and working platform are removed. At the temporary piers, the
support reaction is relieved by jacking the girder, removing the launch bearings, and lowering the
girder in steps until complete release of the jacks.

The vertical displacements are applied sequentially during lowering, in 3060 mm steps at several
piers, to control the transient stresses in the girder. In a three-span bridge, for example, 50 mm
lowering may be applied to one interior pier, then to the other interior pier, then to one abutment
and then to the other abutment, and the entire cycle is repeated until completion of lowering. The
cross-fall may be adjusted by rotating the cross-section during lowering (Zhuravov et al., 1996).

258

Downloaded by [] on [25/09/17]. Copyright ICE Publishing, all rights reserved.


Steel and composite bridges

4.4.2 Launch nose


Launch noses are used frequently during the launch of steel girders to control the launch stresses
in the front region of the girder and to facilitate recovery of the elastic deection of the front
cantilever at landing at the new pier.

A steel girder is much lighter than a PC deck. The launch nose must be lighter than the girder to
control the peak negative bending, and the use of trussed noses is therefore very frequent. The
launch noses for steel girders and PC decks are rarely interchangeable. Braced I-girders or trusses
are used for the shortest noses. Longer noses are designed for higher bending and shear, and each
truss may have a V-section with one bottom chord equipped with launch surfaces and two braced
top chords for enhanced control of the lateral stability of the compression ange under positive
bending. Two V-trusses may be used for the launch of U-girders in combination with a diaphragm
at the nosegirder joint that distributes the axial loads applied by the nose chords (Figure 2.58).

The nose for a steel girder is typically shorter than the nose for a PC deck because of the larger
exural and shear capacity of a steel girder, especially when the girder is completed with a concrete
slab on launch completion. For a conventional continuous beam with an end span 2030%
shorter than the inner spans, a nose with 0.2 , Ln /L , 0.3 is long enough to place the stiffened
girder region of the rst interior support over the lead pier when the nose tip lands at the new pier.
When launching along a curve, this solution also avoids integrative cross-frames for the control of
torsion and warping in the front cantilever.

The hydraulic systems for deection recovery are designed for large vertical movements. A special
lifting frame designed for staged realignment was used to recover the nose deection of the Millau
Viaduct in France. The reduction of negative bending is less of a concern than for a PC bridge, and
the bottom chords of the nose may be inclined or rounded to recover the rst part of the deection
with long-stroke double-acting cylinders, and the remainder by the wedge action of the nose. The
gradient of the bottom chords of the realignment wedge rarely exceeds 810%.

The forward push that friction, launch gradient and wedge action apply to the new launch bearings
is a large part of the vertical support reaction, but the latter is small, and pier overloading is
infrequent. The forward push may overturn the launch bearings, which must be anchored to the
pier cap with stressed tie-downs. Large rotations are also necessary for the launch bearings to
provide realignment by wedge action.

Inclined bottom chords diminish the support action of the launch nose. For the launch stress
analysis of a continuous beam, the nose is modelled as a frame extension with exural and shear
stiffness determined with a separate model of the truss. The inclination of the bottom chords is
modelled by applying to the front support of the model a settlement equal to the distance of the
undeformed bottom chord from the projected launch surface. After determining the support
reactions applied to the nose during the course of launching, the nose is designed by applying
those forces to the truss nose model.

4.4.3 Alternative support systems for the front cantilever


During launching, the peaks of negative bending and shear are reached in the girder region over
the lead pier when the nose is landing at the new pier. Temporary piers have been used in multiple
long-span launches to reduce the launch stresses throughout the length of the bridge. Temporary
front cable-stayed systems have also been used frequently, as a steel girder is much lighter than a

259

Downloaded by [] on [25/09/17]. Copyright ICE Publishing, all rights reserved.


Bridge Launching

PC deck and two stay cables may be used to support the front cantilever. Multiple pull adjust-
ments may also be avoided because of the exibility and the exural and shear capacity of a steel
girder devoid of the concrete slab during launching.

One of the main advantages of a steel girder is easy adaptation. A PC box girder cannot be
strengthened temporarily in the short region where the launch demand exceeds capacity.
However, a steel girder can be strengthened by bolting cover plates, shapes or temporary exten-
sions over the top anges to increase the negative exural capacity in the critical launch region.

A front cable-stayed system is structurally very effective because of the high strength-to-weight
ratio of the prestressing strand and the long lever arm generated by the mast. Alternative
strengthening solutions have been tested to locally increase the negative exural capacity of the
girder.

g The Nizhnegorodsky Bridge in Russia is a steel truss launched on 100 m spans. Two steel
masts were applied to the truss at the fth of the span on either side of the rst interior
support section. The masts were used to deviate temporary cables anchored in the middle of
the rst and second spans. A similar solution was used for the launch of 50 m spans for the
San Roque Viaduct in Spain.
g The Caynarachi Bridge in Peru is a twin I-girder launched with a trussed nose on 50 m
spans. Two vertical trusses braced to each other were temporarily applied to the top anges
of the main girders in the critical launch region.
g The Enlace del Cadagua Bridge in Spain is a twin I-girder launched on 100 m spans.
Segments of twin I-girders were applied to the top anges in the critical launch region to
also solve space issues in the assembly area behind the abutment.

For a temporary strengthening strategy to be a viable option, the girder elements applied to the top
anges should be used in the bridge on launch completion. This would limit the additional cost to
the labour demand for the application and removal of the strengthening systems. When an open
U-girder is launched on long spans, the nal front segment of the girder may be applied upside-
down over the U-girder in the critical launch region (Figure 4.22). A box girder of double depth
would signicantly increase the exural and torsional stiffness and capacity of the critical launch
region. Full-span launching on 150 m spans with a double U-girder has been investigated with
numerical analysis and wind-tunnel testing (Navarro-Manso, 2013).

4.4.4 General launch stresses and instability


The launch stresses in the steel girder are determined with the approach discussed in Section 2.13.
The use of envelopes instead of hundreds of bending and shear diagrams is still advantageous,
but single-line models may be insufcient, and the stress envelopes are often calculated with
three-dimensional grillage models. The envelopes of bending and shear are then checked for
strength and stability of the individual structural members.

An I-girder subject to mainly exural stresses is affected by two modes of instability, both of which
can arise during launching.

g Lateral torsionexure buckling, which is a failure mode that involves lateral movement
and twist of the girder cross-section. A beam loaded in the vertical web plane suddenly
becomes unstable and twists laterally under a critical load smaller than the exural strength.

260

Downloaded by [] on [25/09/17]. Copyright ICE Publishing, all rights reserved.


Steel and composite bridges

Figure 4.22 Double U-girder for full-span launching (figure: Navarro-Manso)

g Local buckling, which may occur in the compression members of cross-frames and lateral
braces, in the web panels of plate girders (along horizontal lines in the case of pure bending,
and along inclined lines in the presence of shear forces), and in the compression anges.

Buckling is the only failure mode of a steel structure that depends primarily on the geometry and
only secondarily on the properties of steel; buckling is also the only failure mode that occurs under
compressive loads.

Most design standards recognise the two forms of instability. Lateral torsionexure buckling is
related to the sway of the entire structural system, and is typically catastrophic in nature. Local
buckling is due to deformations of members between the end nodes (local buckling) or to
cross-section distortion (buckling of anges, webs and stiffeners). Both forms of instability are
related to member slenderness and are characterised by a rapid change in geometry and member
distortion. Inelastic deformation typically follows elastic instability.

4.4.4.1 Lateral torsionflexure buckling


Because of the structural complexity and redundancy of the steel framing system, approximate
values for the critical load for elastic lateral torsionexure buckling are used only during the
preliminary design of cross-frames and lateral braces. Most design standards for I-girder bridges
consider the lateral buckling of individual girders between cross-frames. The classic solution for
the critical moment initiating lateral torsionexure buckling for a simply supported, doubly
symmetric I-girder subject to a uniform moment Mcr about the strong axis was derived under the
assumption of no twist at the ends of the unbraced length (Timoshenko and Gere, 1961):


p
Mcr = Cb EIy GJ 1 + Wr2 (4.14)
L

261

Downloaded by [] on [25/09/17]. Copyright ICE Publishing, all rights reserved.


Bridge Launching

where

p Ekw
Wr = (4.15)
L GJ
represents the warping restraint contribution (Sayed-Ahmed, 2005). In Equation 4.14, L is the
beam span, the shear modulus may be taken as G = E/2.6 for metals, Iy is the lateral moment
of inertia of the girder about the vertical (weak) axis, J is the constant of torsion and kw is the
warping constant. For a doubly symmetric I-girder, the warping constant is kw = Iy(h/2)2, where
h is the vertical distance between the mid-bres of the anges. In-plane bending deformation
increases the lateral stability of the I-girder; however, the Ix /Iy ratio is small in a thin-walled
I-section, and the effects of in-plane bending are typically ignored.

Several design standards specify correction factors Cb for Equation 4.14 to account for the effect of
moment gradients from other types of loading and the location of the load points on the critical
buckling moment of a simply supported I-girder. In most practical cases, however, the equivalent
moment factors are used only to understand the inuence of the different parameters on the critical
moment and as a starting point for the trial-and-error design of cross-frames and lateral braces.

When the I-girders are closely spaced, lateral instability can also affect the steel grillage as a whole.
When the cross-sectional moment of inertia about the horizontal axis is smaller than the moment
of inertia about the vertical axis, the load factor for global buckling is typically adequate for both
cambered and non-cambered geometry. This conclusion, however, is based on the assumption that
the cross-section behaves as a whole about the vertical axis. The concrete slab provides lateral
exural capacity and stiffness on launch completion, and lateral bracing is often designed for
launching only. In the absence of lateral bracing, only cross-frames connect the I-girders, and a
steel grillage with a large span-to-width ratio is susceptible to global lateral torsionexure
buckling during construction. Twin I-girder systems are more susceptible to system buckling, but
multi-girder systems can also fail globally if the girders are closely spaced (Yura et al., 2008).

Closed-form equations have been proposed for the global lateral torsionexure buckling of a
twin I-girder grillage interconnected with diaphragms or cross-frames that stabilise the
compression ange. Under the assumption that the cross-frames are sufciently stiff in their plane
to maintain the same angle of twist for the I-girders, the spacing b of the girders and their lateral
moment of inertia Iy are the principal variables controlling global buckling. The number and
stiffness of the cross-frames have little inuence, and the critical buckling moment for a simply
supported span of length L with both ends restrained from twisting becomes (Yura et al., 2008)

2p p2 E 2 Iy  2 
Mcr = EIy GJ + 2
Iy h + Ix b2 (4.16)
L 4L

where Ix is the moment of inertia of each girder about the strong axis, and all section properties are
those of the individual girder. In many twin I-girder systems, the second term under the radical
dominates. Retaining only the second term under the radical leads to the following estimate of the
critical moment for lateral torsionexure buckling, which shows that the critical moment is
directly proportional to the girder spacing:

p2 bE 
Mcr = Iy Ix (4.17)
L2

262

Downloaded by [] on [25/09/17]. Copyright ICE Publishing, all rights reserved.


Steel and composite bridges

For torsionally braced singly symmetric girders, an effective lateral moment of inertia Iy,eff =
Iy,c + (dt /dc )/Iy,t is substituted for Iy , where Iy,c and Iy,t are the lateral moments of inertia of the
compression and tension ange, respectively, and dc and dt are the distances of the compression
and tension ange mid-bres from the cross-sectional centroid, respectively (Yura, 2001).
Equation 4.16 can also be used for an open U-girder, provided that only one-half of the bottom
ange is used in the calculation of Iy,eff , and for multi-girder systems with obvious adaptations.

These equations are not immediately applicable to the analysis of the front cantilever of a
launched bridge. The tip of the cantilever is neither supported nor prevented from twisting,
bending is not constant along the cantilever, the cross-sections are neither doubly symmetric nor
constant, and local instability within the braced length of the girder often governs design. Finite-
element analysis is therefore necessary in most cases.

Single-line models may be sufcient for rectilinear U-girders stiffened during launching by a
lateral truss at the top ange level. Shear-lag effects may be signicant in the bottom ange and,
if the latter is wide, an effective width is used for analysis. Design standards specify effective width
ratios for a uniformly distributed load for the inner spans and the front cantilever of a continuous
beam. An effective torsional stiffness is also used to account for cross-sectional distortion. The use
of effective cross-sectional properties does not affect the results of analysis signicantly, while their
effects on the structural checks are more marked.

In the most delicate cases (wide U-girders, curved launch alignment, no lateral bracing at the top
ange level, narrow grillages of I-girders equipped with a front cable-stayed system), a three-
dimensional analysis may be necessary. Girders and web stiffeners are modelled with four- or
eight-node shell elements with six degrees of freedom at the nodes to also depict local buckling
modes, the launch rollers are modelled with articulated assemblies of frame elements that generate
the same support reaction at each roll, and the computational effort increases signicantly.
Therefore, three-dimensional models are typically used only to analyse stability of the structural
system, and the general launch stresses are obtained with simpler models.

The three-dimensional models include cross-frames and lateral bracing. The stresses in the bracing
systems due to geometry imperfections and misaligned launch bearings are analysed by imposing
displacements on the support nodes of the model. Displacements at the supports are also applied
to simulate staged lowering of the girder on the permanent bearings at the end of launch. These
analyses may be complex in grillages including several I-girders due to the higher degree of redun-
dancy. Analysis with three-dimensional models is indispensable for curved bridges, as twisting of
curved I-girders is resisted by interaction with adjacent girders.

Additional complication derives from the vertical prole of the girder, which is cambered to reach
the nal vertical alignment after application of the concrete slab. The girder is exible before
achieving composite action and the concrete slab is relatively heavy, and cambering is therefore
signicant, often of some decimetres. The cambers follow the nal span sequence, and, if the
span length is not constant throughout the bridge, the non-launchable cambered prole is offset
from the launch surface at some supports during launching. The hyperstatic effects of the
non-launchable prole are similar to those discussed in relation to the creep deections of a PC
deck. Also in this case, the girder is analysed as a rectilinear one, and vertical displacements
corresponding to the distance of the cambered prole from the theoretical launch surface are
applied to the support nodes of the model.

263

Downloaded by [] on [25/09/17]. Copyright ICE Publishing, all rights reserved.


Bridge Launching

Figure 4.23 Lateral torsionflexure buckling of the front cantilever

Lateral torsionexure buckling is a prime concern in the incremental launching construction of


steel girders. The girder is often launched with a short launch nose, and the long front cantilever is
prone to lateral buckling (Figure 4.23). Lateral torsionexure buckling, however, is not necess-
arily more critical than local forms of instability. For example, when a cable-stayed system is used
to support the front cantilever during launching, the longitudinal component of the pull in the stay
cables induces axial compression in the steel girder, and local instability may have a lower buckling
factor than lateral torsionexure instability.

If the compressive force in a member is large enough, the transverse stiffness goes to zero, the stiff-
ness matrix becomes singular and the member is said to have buckled. Linear buckling analysis
seeks the instability modes of a structure due to the P-delta effect under a specied set of loads
through the solution of the generalised eigenvalue problem. Each eigenvalueeigenvector pair is
called a buckling mode of the structure. The eigenvalue is termed the buckling factor, and it is the
scale factor that must multiply the loads to cause buckling in the given mode. The criticality of a
buckling mode, therefore, is measured by its buckling factor.

The buckling modes of a structure depend upon the load. There is no set of natural buckling
modes like there is for the natural vibration modes. Since bridge launching involves the
complication of deck movement, the buckling modes and the corresponding buckling factors are
calculated for different load cases and different support congurations of the deck. The structural
analysis programs identify the buckling modes with increasing buckling factors. Sometimes, the
rst modes are global modes of the structural system such as lateral torsionexure buckling of
the front cantilever or out-of-plane buckling of a temporary pier, and the analysis progresses
toward more and more localised modes with shorter effective length. Other times, however, some
local modes have a lower buckling factor than the global system modes.

264

Downloaded by [] on [25/09/17]. Copyright ICE Publishing, all rights reserved.


Steel and composite bridges

Figure 4.24 Local buckling of an unstiffened web panel

The analysis of lateral torsionexure buckling does not require the use of shell elements. If the
I-girders are modelled with four- or eight-node shell elements with six degrees of freedom at the
nodes, if the model includes web stiffeners, and if meshing is designed for mesh-independent
results, buckling analysis can also identify local modes in webs and anges (Figure 4.24). Local
buckling may be critical at the root of the front cantilever, especially when a cable-stayed system
is used to support the cantilever and the mast applies a signicant vertical load to the top ange of
the girder. The launch rollers must be modelled accurately to ensure realistic dispersal of the
support reactions within the webs and to avoid over-restraining the bottom anges.

Investigating many buckling modes with detailed structural models involves a long computation
time. In many cases, only lateral torsionexure buckling is analysed numerically, and local
buckling is checked with the equations specied by the design standards in relation to the
cross-sectional geometry and the effective length and the end restraint of members. Integrative
load factors are sometimes specied to allow for eccentricity, distortion and fabrication-induced
residual stresses.

Some types of cross-section require the numerical analysis of local buckling as well. This is
often the case for the I-girders of long-span composite bridges, because the midspan sections are
different from the pier sections, the girders are tall, and the cross-sections are typically Class 4
slender sections. The use of a front cable-stayed system further complicates matters because of the
axial compression generated by the longitudinal component of the pull in the stay cables.
Figure 4.25 shows web buckling in the central region of the rear span due to self-weight bending
and axial compression.

265

Downloaded by [] on [25/09/17]. Copyright ICE Publishing, all rights reserved.


Bridge Launching

Figure 4.25 Local web buckling due to self-weight bending and axial compression

Buckling is analysed with the stiffness matrix written for the deformed geometry under the loads
that generate buckling. Large-displacement analysis is recommended for the long front cantilevers
of launched bridges. Geometry imperfections may be surveyed and applied to the model with
equivalent lateral loads or by modifying the node coordinates. The buckling factors of twin
I-girders and multi-girder grillages are sensitive to geometry imperfections because, in one of the
many structural congurations for launching, the imperfections may have the same shape as a
buckling mode (Rosignoli, 2013). Since lateral torsionexure buckling is a ULS condition, the
buckling factors are calculated for the factored load combinations specied by the design
standards for temporary construction conditions. Self-weight, lateral wind and some geometry
imperfections are typically included in the analysis.

The resistance factors specied by the design standards for the different modes of instability are
applied to the buckling factors determined with factored loads to calculate the governing
capacity-to-demand ratio. Eurocode 3 (EN 1993-1-1:2005: UNI, 2005a) determines the vertical
exural capacity for lateral buckling based on a section modulus related to the cross-sectional
slenderness, the material resistance factor, and a reduction factor that depends on an imperfection
factor and on the ratio of the cross-sectional exural capacity to the elastic critical moment for
lateral torsionexure buckling resulting from analysis. The imperfection factor is determined
based on four buckling curves specied for different types of cross-sections. AASHTO (2012)
species f = 0.8 for web crippling and f = 0.9 for axial compression, but does not provide a specic
resistance factor for lateral torsionexure buckling. In the light of the catastrophic consequences of
lateral buckling and of the uncertainties in the analysis, a resistance factor f 0.5 is recommended
for the front cantilever of a launched bridge.

In twin I-girders and multi-girder grillages, lateral torsionexure buckling is controlled by the
cross-frames, which avoid relative movements between anges and control distortion, and by the
lateral braces between anges, which increase the constant of torsion of the structural system, resist
the torsional ow of forces, and distribute the localised stiffening action of the cross-frames along
the girders. In the U-girders, lateral bracing is applied between the top anges, and cross-frames
are added to increase the capability of the cross-section to resist distortion. When the bridge is

266

Downloaded by [] on [25/09/17]. Copyright ICE Publishing, all rights reserved.


Steel and composite bridges

curved in plan, the vertical loads cause primary twisting that requires sufcient torsional strength
for the system. Single and twin U-girders or closed box girders are typically preferred to I-girders
in curved bridges because of their higher torsional strength and stiffness.

I-girders should be restrained at the supports against twisting about the longitudinal axis. Direct
restraint with cross-frames is impossible during launching due to the movement of the grillage.
Indirect restraint is sought by keeping the bottom girder anges guided laterally at the launch
bearings and by enhancing the interaction between the components of the structural system.
Several design standards provide instructions for the optimum sizing of the lateral bracing between
anges. Eurocode 3 (EN 1993-1-1:2005: UNI, 2005a) takes into account the imperfections of
welded girders and provides information on the lateral torsionexure stability of symmetric and
non-symmetric sections.

4.4.4.2 Local buckling


Most design standards cover the local buckling of web panels. In a non-launched bridge, this form
of instability is rarely a potentially critical situation such as the out-of-plane buckling of slender
compression members or the lateral torsionexure buckling of the entire structural system, since
anges and vertical stiffeners restrain the edges of the web panel, and membrane stresses develop
a natural stiffening action in the panel and resist the growth of out-of-plane deformations.

After buckling, the web plates develop a strength reserve due to the elastic work of membrane
stresses. In the pre-critical exural state, the girder deforms keeping the sections plane. At buck-
ling, the axial stress redistributes toward the stable zones of the web panel, the slope of the tension
diagonal changes, and a new equilibrium based on higher stresses arises.

Years ago, instability of web panels was analysed with manual methods. When the plate began to
assume out-of-plane deformations, the loading was interrupted and the value of this critical
stress was calculated. The present means of testing and calculation conrm that, beyond this
critical stress, the growth rate of out-of-plane deformation does not increase, but tends to decrease
until rst yielding occurs. Only at this point does the deformation begin to increase more quickly
than the load, and the plate collapses. The critical stress identies the transition between two
different states of elastic equilibrium, and does not constitute the limit value of the carrying
capacity of the panel, which is much higher.

The ratio of the ultimate stress to the critical stress increases with the plate slenderness, varies
between 3 and 8, and for the web slenderness of composite bridges is in the range 56.
However, the web panels are not perfectly at and are often affected by residual welding stresses.
Test models showed that, because of this, the webs use a portion of the post-critical domain at the
SLS, and the margin from the ultimate stress is therefore smaller. The post-critical behaviour of a
non-stiffened plate is also different from that of a stiffened web panel.

Some design standards base the analysis and the strength checks of steel and composite sections on
the concept of effective breadth. This concept is applied to webs, anges and the bottom ange
zones of U-girders between longitudinal stiffeners. Recent design standards such as Eurocode 3
(EN 1993-1-1:2005: UNI, 2005a) and Eurocode 4 (EN 1994-1-1:2004: UNI, 2005b) offer signicant
advances and introduce the nonlinear instability theory. This approach permits accounting for the
post-critical behaviour of slender plates subject to normal and tangential stresses, according to
resistance factors from the ULS deriving from the most recent research.

267

Downloaded by [] on [25/09/17]. Copyright ICE Publishing, all rights reserved.


Bridge Launching

In the ideal case of a beam suffering no local buckling of webs or anges, the cross-section can
develop the full plastic moment, possesses a high rotation capacity, and is classied as a Class 1
compact section. Class 2 compact sections can also develop a full plastic moment, but have less
rotation capacity. When an I-section has a slender web or compression ange, local buckling may
occur at the web or ange prior to developing the full plastic moment. If local buckling occurs
after developing the rst-yield moment, the section is classied as a Class 3 non-compact section.
Finally, if local buckling occurs prior to reaching the rst-yield moment, the section is classied as
a Class 4 slender section.

Most launched bridges have Class 3 or Class 4 sections. The exural capacity of a Class 4 section is
based on the mechanical properties of an effective cross-section that is determined based on the
concept of effective breadth. The design standards provide equations for determining the effective
breadth of the members of I-girders based on the outstand-to-thickness ratio of the compression
ange and the depth-to-thickness ratio of the web panel. The exural capacity Mr of a Class 3
section depends on the critical moment Mcr for lateral torsionexure buckling and on the rst-yield
moment My of the cross-section. For example, according to CAN/CSA-S16.1-94 (CSA, 1994), the
exural capacity of a Class 3 section is Mr = fMcr when the critical moment is Mcr 0.67My, and
 
My
Mr = 1.15fMy 1 0.28 fMy (4.18)
Mcr
when the critical moment is Mcr . 0.67My; f is the material resistance factor.

4.4.5 Local launch stresses and instability


Erection by incremental launching is particularly suited to U-girders and twin I-girders, which
both require two support lines at each pier. This facilitates the launch, reduces the cost of the
launch bearings and simplies the analysis of the launch stresses. Launching three or four braced
I-girders is more expensive, and misaligned launch bearings may overload the cross-frames
between the girders. Four-girder grillages may be launched in one operation, or two twin I-girders
may be launched independently and connected on launch completion. When multi-girder grillages
are launched in one operation, the central pair of girders is braced laterally, and the cross-frames
stabilise the outer girders (Figure 4.2).

Complex states of stress arise in the girder regions immediately above the launch bearings.

g Self-weight bending, the hyperstatic effects of the non-launchable prole of a cambered


girder, geometry imperfections and thermal gradients generate axial and tangential stresses
in the longitudinal plane.
g Torsion and distortion due to misaligned launch bearings and geometry imperfections
generate axial and tangential stresses in the longitudinal and transverse planes.
g Dispersal of the support reactions within the webs generates local vertical axial stress.
g Transverse eccentricity of the support reactions and planarity tolerances in the bottom
ange and the launch rollers generate transverse bending in the unstiffened bottom web-
ange node.
g Launch friction and the local launch gradient generate local longitudinal axial stress that
compromises the conservation of plane sections.

Most of these stresses may be signicant, and the state of stress is checked with biaxial or triaxial
strength criteria.

268

Downloaded by [] on [25/09/17]. Copyright ICE Publishing, all rights reserved.


Steel and composite bridges

4.4.5.1 Vertical axial stress


On launch completion, the girder is supported at the nal pier sections, where vertical web stiffeners
distribute the support reactions to the web panels and prevent local instability. The nal distribution
of stiffer and lighter sections enhances exural and shear capacity, and the girder is designed to resist
the weight of concrete slab and construction equipment prior to the onset of composite action.

During launching, the support sections are devoid of stiffeners. Although the girder is much
lighter than the nal composite deck, the strength and stability of unstiffened web panels subject
to patch loading may be critical. Dispersal of the support reactions within the webs may require
triangular corner stiffeners throughout the length of the bridge (Figure 2.50). Corner stiffeners
increase the torsional stiffness of the webange node and provide local beam action and direct
load transfer between the vertical web stiffeners that enhances the longitudinal dispersal of
support reactions (Mato and Santos, 2007).

The vertical compressive stress in an unstiffened web panel due to a support reaction applied to the
bottom ange is calculated under the assumption that the load is dispersed uniformly within the
web. Load dispersal may be assumed to take place at an angle of 608 from the edge of the patch
load through the thickness of the bottom ange, and then at an angle of 458 through the web panel
itself (Figure 4.26).The exibility of the launch bearings and the longitudinal exural stiffness of
the bottom ange help in making the patch loading uniform. In the most delicate cases (sliding
bearings with long rocking frames, angle breaks in the bottom ange, inclined webs), nite-
element analysis of the interaction between launch bearings and the steel girder may be necessary.

BS 5400 (BSI, 1982) allows the effects of in-plane patch loading to be disregarded when the vertical
compressive stress sz in the web plate is (in megapascals)

tw 355
sz 3fy  (4.19)
Lw hw fy

where Lw is the length of patch loading measured on the longitudinal web edge, tw is the web plate
thickness, hw is the net depth of the web panel measured clear between anges and fy is the nominal

Figure 4.26 Load dispersal within an unstiffened web

Lw

45 60

Lb

269

Downloaded by [] on [25/09/17]. Copyright ICE Publishing, all rights reserved.


Bridge Launching

Figure 4.27 Local failure modes of an unstiffened web

Local yielding Local buckling Panel buckling

yield strength of steel. In most cases, the vertical compressive stress due to the launch support
reactions is higher than this threshold.

An unstiffened web panel subject to a concentrated support reaction applied through the bottom
ange is affected by three failure modes, which depend on the load intensity and on the slenderness
of the web panel (Figure 4.27).

g The rst mode is caused by web yielding immediately above the load, followed by the
plastic deformation of the bottom ange. This mode is called local yielding.
g The second mode is triggered by localised buckling in the lower portion of the web panel,
for a depth about 50 times the plate thickness, followed by web yielding and the onset of a
plastic mechanism in the bottom ange (Bernard-Gely and Calgaro, 1994). This mode is
called local buckling.
g The third mode is the general buckling of the web panel extended to most of its depth. This
mode is assessed with the equations specied by the design standards for panel buckling
under shear and bending.

The cross-sectional strength Rr for a support reaction applied orthogonally to the web plate
through the bottom ange is the lower value of the local yielding strength Ry and the local buck-
ling strength Rb . According to Eurocode 3 (EN 1993-1-1:2005: UNI, 2005a), the local yielding
strength of an I- or U-section is
Ry = 0.9(Lb + Ly )tw fy (4.20)
where Lb is the length of the rigid patch loading measured on the launch surface (lower edge of
bottom ange) and Ly is

 
bb sx 2
Ly = 2tb 1 1.1 (4.21)
tw fy

where sx is the longitudinal axial stress in the bottom ange, and bb and tb are, respectively, the
width and thickness of the bottom ange, with the limitation bb /tb 25. Always according to
Eurocode 3 (EN 1993-1-1:2005: UNI, 2005a), the local buckling strength of the web panel is

tb t L
 +3 w b
t tb hw
Rb = 0.5t2w Es fy w (4.22)
f

270

Downloaded by [] on [25/09/17]. Copyright ICE Publishing, all rights reserved.


Steel and composite bridges

where f is the resistance factor of material, with the limitation Lb /hw 0.2. The design value Rd of
the support reaction applied to one girder is determined based on the maximum support reaction
from analysis, the expected misalignment of the launch bearings, and the expected geometry irre-
gularities in the bottom anges due to fabrication and assembly tolerances. A 30% increase in the
maximum support reaction from analysis is sometimes sufcient for I-girder grillages
(Chemerinski et al., 1996), provided that stringent support and geometry tolerances are specied
and actually respected. Higher values should be considered for U-girders because of the higher
torsional and distortional stiffness of the cross-section. Based on the combined load factor g 1
specied by the applicable design standards for the load combination and the limit state under
consideration, and on the resistance factor f 1 specied for the governing failure mode and the
limit state under consideration, it must be gRd fRr .

The exural capacity Mr of the cross-section is determined as the lower value of the plastic
moment and of the critical moment for buckling. Setting Md as the design value of the bending
moment in the cross-section, Eurocode 3 (EN 1993-1-1:2005: UNI, 2005a) species an additional
stability check to account for the possibility of the simultaneous onset of both buckling modes,
gRd gMd
+ 1.5 (4.23)
Rb Mr
and, of course, gMd fMr . Finally, the state of stress in the webs is checked with conventional
criteria for biaxial strength.

4.4.5.2 Longitudinal axial stress


The launch surface is rarely horizontal, and the support reactions are not orthogonal to the launch
surface. The girder is almost always launched along an inclined plane, and the vertical component
of the contact forces (i.e. the design value Rd of the support reaction applied to one girder) is not
orthogonal to the bottom ange. The frictional resistance of the launch bearings rotates the
resultant vector and generates additional longitudinal components. These systems of forces are
easy to determine when the local launch gradient and the design value of the support reaction are
known. The longitudinal stress may be high before dispersal within the cross-section, and it adds
to the longitudinal compressive stress generated by negative bending in the continuous beam.

The ULS longitudinal axial stress in the bottom ange also depends on the type of girder. For an
I-girder, the stressstrain relation is bilinear. In the absence of instability, Class 1 and Class 2
sections according to Eurocode 3 (EN 1993-1-1:2005: UNI, 2005a), collapse is not reached when
the rst yield occurs at a moment My but when the moment applied to the cross-section reaches the
plastic moment capacity Mp . The ratio Mp /My depends on the cross-sectional geometry and the
type of steel, and for Class 1 and Class 2 sections one may take Mp /My 1.15. Class 3 sections
have Mr = My , since instability prevents the post-yielding domain from being exploited, while the
exural strength of Class 4 sections is governed by instability.

The ULS behaviour of a U-girder subject to negative bending and shear is more complex. Because
the cross-sections do not remain plane, simple beam theory is inadequate to predict ange stress
distribution. If the webs are stocky and only the bottom ange is slender, rst yielding occurs in
the outer ange panels where shear lag amplies axial and tangential stresses. Since the ange
plate and stiffeners are exible, the stresses are magnied at the plate surface due to out-of-plane
buckling, and both My and Mp may be smaller than their theoretical values. Two assumptions
may be made:

271

Downloaded by [] on [25/09/17]. Copyright ICE Publishing, all rights reserved.


Bridge Launching

g the load-carrying capacity of the U-girder is exhausted when rst yielding occurs at the
outer panel of the bottom ange
g the bottom ange has the capacity to redistribute the axial stress across its breadth so as to
force yielding into the less stressed central region of the panel.

In reality, the load-carrying capacity depends on whether the outer panels continue to carry their
share of the load as redistribution takes place or they rapidly shed load due to local buckling. In
order to understand the ULS behaviour of the compression ange of a U-girder, therefore, the
behaviour of each component must be predicted up to and beyond collapse, including the effect
of boundary conditions. This analysis includes the effects of shear lag, restrained warping,
cross-sectional distortion and geometry imperfections.

Many design standards cover the effects of shear lag. Among the parameters considered are ange
plate orthotropy, ratio of ange to web area, cross-sectional geometry, aspect ratio of the anges,
position of point loads, load combinations, and support conditions. Some standards specify
effective width ratios for uniformly distributed load for the interior spans of the continuous beam
and for the front cantilever. BS 5400 (BSI, 1992) also species the effective web thickness to be
used for exural stress analysis of beams devoid of longitudinal stiffeners.

With the exception of circular sections and square thin-walled sections, warping of the cross-
section occurs when the section twists (Belluzzi, 1988). If warping is resisted, restrained warping
stresses are induced in the section. Some design standards specify equations to predict the peak
warping stresses in the box girders. When the girder is curved in plan, the axial and tangential
stresses due to non-uniform torsion and restrained warping may be signicant. AASHTO
(2003) species reductions in the SLS axial stress limits in the ange plates based on the geometric
characteristics of the cross-section and the inuence of non-uniform torsion. Eccentric loads also
twist and distort the cross-section between internal diaphragms. Although distortion is small in terms
of displacement, it may cause signicant transverse bending and longitudinal warping stresses
(Section 3.4.2). Several design standards provide equations for the analysis of these stresses as well.

The ULS behaviour of the bottom compression ange of a U-girder can be predicted with elasto-
plastic large-deection analysis. The effects of residual stresses and the initial lack of atness can
be included in the analysis, although it is difcult to measure the geometry imperfections. Residual
stresses have a predominant weakening effect on practical plates in compression. Analyses showed
that geometry imperfections can reduce the plate strength signicantly, while cylindrical out-of-
plane distortion has little weakening effect. Tests showed that the transverse llet welds between
the bottom plate and diaphragms do not reduce the plate strength, whereas stepping at a butt weld,
particularly where a thinner plate is connected, does reduce the strength.

Xanthakos (1994) suggested the following ultimate axial stress for uniaxial compression:
 
fy
sult = fy 1 , kfy (4.24)
8kscr

where fy is the yield strength, scr is the critical buckling stress and k is the ratio of the average axial
stress at rst surface yielding to the yield strength. The factor k is obtained from large-deection
elastic analysis, and varies with the width-to-thickness ratio, the assumed geometry imperfections,
and the degree of end restraint.

272

Downloaded by [] on [25/09/17]. Copyright ICE Publishing, all rights reserved.


Steel and composite bridges

For the above reasons, the longitudinal axial stress in the support regions of a continuous U-girder
differs from Naviers stress distribution and reaches higher values in the bottom ange. Some of
the causes of stress nonlinearity depend on the tributary girder length on the launch bearings, and
their effects are therefore a maximum in the lead support region of the girder at nose landing on
the new pier.

Since the end span of the bridge is typically shorter than the interior spans, the length of the launch
nose may be chosen so as to locate the stiffened region of the rst interior support over the lead
pier when the nose tip lands on the new pier. If the stresses in the long front cantilever are
excessive, the nose may be lengthened and the stiffened region at the rst interior support may
be extended backward with additional cross-frames and web stiffeners. As an alternative, a front
cable-stayed system may be used to control the excess of negative bending.

4.4.5.3 Stress checks


When passing above the launch bearings, the lower region of the webs of I- and U-girders is
subject to a triaxial state of stress. The lateral guides keep the support reactions aligned with the
webs, transverse bending is relatively low, the state of stress can often be assumed as a plane one,
and the HuberMises ideal stress can be directly compared with the factored yield point of steel for
the limit state under consideration:

sid = s2x + s2z sx sz + 3t2xz ffy (4.25)

The longitudinal axial stress sx due to negative bending and the vertical axial stress sz due to the
dispersal of the support reaction within the web have the same sign and decrease along the web
depth. Even if the tangential stress txz is smaller at the webange nodes than at the middle of the
web panel depth, the highest values of sid are typically reached near the weld of the web to the
bottom ange, which must be checked accurately during fabrication. Design standards such as
AASHTO (2012) specify the use of the same resistance factors for welds and base metal, and the
use of f = 1.0 for tension or compression normal to the effective weld area.

High transverse bending may arise in the unstiffened bottom webange node when the support
reactions are not aligned with the webs, when the contact surface of cast-iron rolls or rollers is not
horizontal, or when the bottom ange is not horizontal due to fabrication tolerances. For these
reasons, the transverse width of the launch bearings should be kept to a minimum. Weld hammer-
ing methods such as HiFIT have proven to be effective and inexpensive treatment methods to
increase the service life of welded steel structures. A hardened pin hammers the weld toe with
an adjustable frequency of 160300 Hz. The weld toe is deformed plastically, and the induced
residual compressive stress prevents crack propagation on the surface (Rosignoli, 2013).

The HuberMises ideal stress in the bottom ange of a U-girder is



sid = s2x + 3t2xy ffy (4.26)

The distribution of the longitudinal axial stress sx depends on the vertical and lateral bending,
restrained warping and shear lag. The torsional tangential stress txz can be high when the girder
is launched along a curve. The combined effects of torsion, restrained warping, shear lag and
exure are difcult to determine, and the worst case of any of these may be superimposed to obtain
a conservative estimate.

273

Downloaded by [] on [25/09/17]. Copyright ICE Publishing, all rights reserved.


Bridge Launching

Eurocode 3 (EN 1993-1-1:2005: UNI, 2005a) species similar criteria for the ULS assessment of
webs subject to support reactions and design shear forces below and above 50% of the plastic
shear capacity of the support section. In the rst case, the strength criterion is
      
sx,d 2 sz,d 2 sx,d sz,d
+ 1 (4.27)
ffy ffy ffy ffy
where sx,d and sz,d are the factored design values of the local stresses, and the resistance factor
applied to the yield strength of steel depends on the limit state under consideration. Similar
equations are specied for design shear forces higher than 50% of the plastic shear capacity and
when the exural strength is determined by accounting for the plastic stress redistribution within
the cross-section.

4.5. Launch of trusses


Several steel trusses have been built by incremental launching. Compared with the launch of a twin
I-girder, several factors complicate the launch of a truss.

g Steel trusses are used on spans longer than those achievable with I-girders of reasonable
depth, and long spans complicate launch. Short-span trusses are rarely used for bridge
structures because they require more hand welding and involve higher fabrication and
erection costs than braced I-girders. Modern castellated trusses offer lower fabrication
costs, simpler welding with prole-tracking equipment and no gusset plates at the eld
splices. In spite of this, however, they are still more expensive than braced I-girders.
g When the truss carries a concrete slab, the slab is designed for minimal weight. Live loads
are proportionally less demanding on longer spans, dynamic allowance is lower as well,
self-weight is a larger proportion of the design load, and the strength margins of a truss
during the launch are smaller. The strength margins may be more favourable when the truss
is designed for roadway trafc on the top chord and railway trafc on the bottom chord
(Malite et al., 2000).
g Long-span trusses are designed for permanent load distribution, where some diagonals
resist compressive forces and others resist tensile forces. The load in the diagonals varies in
relation to the presence or absence of live loading; the stress variations, however, are much
smaller than the self-weight stress reversals during launching.
g The bottom chords are designed as axial members but act as runway beams during
launching. Migration of the support reactions generates local exural and shear stresses in
the bottom chords that may lead to oversized plates and eld splices, and that often suggest
monitoring (Malite et al., 2000).
g The diagonals converge into the chords from different directions. Geometry imperfections
cause local stresses during launching and the stress distribution within the connections may
be particularly complex, so solid three-dimensional nite element analysis is often warranted.

When the bottom chord has short panels, paired launch bogies may be used to support the truss in
two adjacent panels. The launch bogies may be placed onto interconnected hydraulic jacks to
generate a hydraulic exural hinge. This improves the local load distribution, but does not reduce
the general off-node eccentricity of the support reaction.

A castellated box chord can be supported under the webs during launching, provided that the
bottom surface of the plates is machined with precision. When an I-beam is used for the bottom
chord, a rectangular launch rail is welded to the bottom ange under the web to facilitate dispersal

274

Downloaded by [] on [25/09/17]. Copyright ICE Publishing, all rights reserved.


Steel and composite bridges

of the support reaction within the web and for lateral load transfer by acting against the anges of
the cast-iron rolls of the launch bogies. When bolted, the eld splices in the bottom chord are
designed to avoid interference with the launch rollers. Details developed for the launch systems
of the self-launching gantries may be easily recongured for a launched truss (Rosignoli, 2013).

4.6. Concrete slab


Orthotropic steel decks and concrete slabs are the two alternatives for the deck platform of steel
girders and trusses. Orthotropic decks are mostly used on long-span bridges because of high
fabrication and maintenance costs. A concrete slab is used as the decking system in most cases,
and the slab is fabricated on launch completion of the steel girder for weight reasons. Three
construction methods are typically used for the concrete slab:

g crane assembly of full-depth precast panels and connection to the steel girder with closure
pours and stud pockets
g crane assembly of left-in-place forming systems and slab completion with in-place concrete
pours
g in-place segmental casting with a motorised forming carriage.

Incremental launching of the concrete slab over the steel girders is a valid alternative but is not yet
well known within the bridge industry.

The steel girder has a stable elastic behaviour while the concrete slab is affected by time-dependent
phenomena, and stress redistribution occurs within the composite deck over time. Without pre-
stressing, cracking of the concrete slab is practically unavoidable. The most critical tensile stresses
are typically induced in the concrete slab during construction. Even adopting staggered sequences
of segmental slab casting, the self-weight and superimposed dead load generate tensile stresses in the
support regions of the continuous beam. The live loads increase negative bending, but permanent
tensile stress is often present even in the midspan regions due to drying and thermal shrinkage, creep,
and the low compressive stress generated by composite action. Longitudinal cracks are also frequent
in the slab regions above the steel girders in the absence of transverse prestressing.

Slab cracking affects the concrete cover to top reinforcement. A cracked slab is subject to the
adverse effects of water and de-icing salts, and this condition is particularly severe when the
wearing surface constitutes an integral part of the concrete slab. When an asphalt layer is used
as a protection, water-tightness is hard to achieve even with high-quality membranes between the
asphalt and the concrete slab. Polyester concrete overlays are typically applied without a mem-
brane. Several design standards stipulate increased concrete cover to top reinforcement to
provide for cracking and deck wear, but this is inconsistent with the practice of placing the
reinforcement near the surface for crack control. However, when the reinforced-concrete slab is
designed correctly, crack widths of about 0.3 mm are often unable to cause corrosion of reinforce-
ment steel.

Slab cracking is mainly caused by exural stresses and shrinkage of concrete, and several design
standards provide instructions in both directions. A minimum reinforcement ratio in the slab
regions where the longitudinal tensile stress exceeds the effective tensile strength of concrete is
often required. The minimum reinforcement ratio is typically set to 1% of the gross cross-sectional
area of the slab, and is increased to 1.31.5% in the support regions of the continuous beam. In
these regions, the SLS tensile stress in the reinforcement is limited to 6070% of the yield point of

275

Downloaded by [] on [25/09/17]. Copyright ICE Publishing, all rights reserved.


Bridge Launching

steel, to further control crack width. Additional provisions govern the quality and compactness of
the concrete, the water-to-cement ratio and the use of admixtures, as all of these parameters
inuence the tensile strength and permeability of concrete.

The crack width mainly depends on the distance to the nearest reinforcing bar placed perpendicu-
lar to the plane of crack, the distance to the neutral axis and the average surface strain at the point
considered. The crack width is controlled by the rebar spacing and diameter and by reduced SLS
tensile stress in the reinforcement, and is checked against the crack width allowed by the design
standards for the environmental conditions of the site by accounting for the effects of the tension
stiffening of the concrete between cracks. According to Eurocode 4 (EN 1994-2:2005: UNI, 2006),
the tensile stress ss,0 generated in the reinforcement by negative bending acting on the composite
section is calculated with the composite section consisting of steel girders and longitudinal
slab reinforcement. In composite decks where the concrete slab is assumed to be cracked and not
prestressed by tendons, the additional tensile stress due to the tension stiffening of concrete
between cracks is

0.4ftm
Dss = (4.28)
as rs

where ftm is the mean tensile strength of concrete. The reinforcement ratio is given by rs = Ar /Ac,eff ,
where Ac,eff is the cross-sectional area of the concrete slab within the effective width and Ar is the
total area of the longitudinal reinforcement within Ac,eff . The coefcient as = AI/AsIs depends on
the area A and the moment of inertia I of the effective composite section, neglecting concrete in
tension; As and Is are the corresponding properties of the structural steel section. The tensile stress
in reinforcement,

ss = ss,0 + Dss (4.29)

is compared with the limitations set by the design standards or is directly used for the analysis of
the crack width. The construction sequence of the composite deck is considered by calculating the
stresses caused by loads applied separately to the steel girder alone, to the short-term composite
section and to the long-term composite section. The cross-sectional properties are evaluated based
on the short-term modular ratio n0 = Es /Ec . The time-dependent effects are taken into consider-
ation by increasing the short-term modular ratio

n = n0[1 + k1k2(t, t0 )] (4.30)

where k2(t, t0 ) is the concrete creep coefcient and k1 depends on the type of action. According to
Eurocode 4 (EN 1994-1-1:2004: UNI, 2005b), k1 is equal to 0 for short-term actions, 0.55 for
primary and secondary effects of shrinkage, 1.1 for permanent loads and 1.5 for slab prestressing
through imposed deformations.

4.6.1 Full-depth precasting


Full-depth precast deck panels have been used since the early 1960s. Their use has become the
standard solution for the composite decks of cable-stayed bridges. A rectangular open grid of edge
girders, oorbeams and substringers is used to support rectangular full-depth panels of trans-
portable dimensions (Figure 4.28). Precast deck panels are also used in multi-girder grillages and
girdersubstringer systems that provide multiple support lines for the concrete slab. The research
behind the full-depth precasting of deck panels is mostly driven by minimal impact requirements
during the rapid replacement of existing bridges.

276

Downloaded by [] on [25/09/17]. Copyright ICE Publishing, all rights reserved.


Steel and composite bridges

Figure 4.28 Precast deck panels of the Arthur Ravenel Bridge in the USA. (Photo: Buckland and
TaylorStarmer)

Increasing volumes of trafc and congestion of existing infrastructure require bridges that can be
constructed rapidly and with a minimum impact on trafc. Steel girders offer high quality and
accurate geometry from industrial fabrication processes, and incremental launching offers rapid
erection of steel girders with no impact on trafc. Casting the concrete slab in-place becomes the
bottleneck in terms of construction rapidity and the impact on trafc (Stanton et al., 2005). Precast
concrete construction provides a promising alternative to in-place casting, with signicant
reductions in construction time, decreased impact on the travelling public, and similar durability
and structural performance (NCHRP, 2008).

Full-depth slab precasting offers the same advantages as the segmental precasting of PC bridges,
such as dimensional stability due to long curing, accurate dimensions, and high quality resulting
from industrialised casting processes. Precast concrete members are more durable and more uni-
formly constructed than their cast-in-place counterparts because of the controlled fabrication
environment and stringent quality control in precasting facilities. A study on bridge durability
found that a lower percentage of precast PC bridges were structurally decient than cast-in-place
PC bridges of a similar age and span length (Dunker and Rabbat, 1992). Precast concrete
components are advantageous for bridges being constructed over water, wetlands, and sensitive sites
where environmental concerns and regulations discourage the use of fresh concrete. Precast concrete
systems also require signicantly less formwork than cast-in-place concrete. This provides an
additional advantage for situations in which clearance is limited and sufcient room is not available
for formwork extending below the deck (Stanton et al., 2005).

Full-depth precast deck panels span the width of the bridge and are placed adjacent to each other.
They are typically cast end-to-end on a long-line prestressing bed, and may be pre-tensioned
transversely and post-tensioned longitudinally. Transverse prestressing allows for thinner panels,

277

Downloaded by [] on [25/09/17]. Copyright ICE Publishing, all rights reserved.


Bridge Launching

provides better crack control, and facilitates handling, transportation and erection. The panels are
supported on the steel girders with exible shims or leveling bolts for geometry adjustment and
compensation for geometry irregularities in girders and substringers. The solution of continuity
over the anges is lled with grout, to provide a uniform bearing surface for the panels, and the
leveling bolts are backed out or removed, to avoid stress concentration.

The main disadvantage of slab precasting is the small segment width deriving from transport
requirements. This increases the number of construction joints in the slab, with related costs and
durability issues. Non-reinforced female-to-female grouted joints or different types of reinforced-
concrete joints are used between adjacent panels. Longitudinal post-tensioning may be used to
keep the transverse joints in compression. Shear studs are welded to the supporting girders within
pockets in the panels, to create a composite deck. Full-depth panels require little or no formwork,
and eliminate the cure time needed for cast-in-place decks.

The weak point of precast deck panels is the great number of joints and stud pockets. Leaking and
spalling of the joints have been historically associated with this construction method, although
improved joint details are likely to reduce these problems. Signicant effort is also required to level
the panels and to grout the joints between the panels and the haunches under the panels. Special
proprietary materials have been proposed for grouting and the in-place casting of connections.
These materials have high bond when cast against hardened concrete and a short bond
development length with embedded reinforcement, which results in narrow in-place stitches that
are stronger, tougher and more durable than the adjoining deck panels. The advantages include
reduced joint size and complexity, improved durability and continuity, fast construction and low
maintenance.

Powerful ground cranes or crane-trolleys running on the steel girder are necessary to handle wide
full-depth precast deck panels, and this increases the nal cost of the solution. For these reasons,
segmental deck precasting is used mostly for cable-stayed bridges, special accelerated-construction
applications, and in cold weather to minimise eld operations involving fresh concrete (Monov
and Seliverstov, 1997).

4.6.2 In-place casting


Full-depth precast deck panels are a cost-effective solution for multi-girder grillages and girder
substringer systems, where the short transverse spacing of the support lines of the concrete slab
allows the use of 200250 mm panels. Full-depth precasting is also the standard solution for the
composite grillages of cable-stayed bridges, where oorbeams between the edge girders provide
transverse support lines for the concrete slab.

Handling precast deck panels thicker than 200250 mm is difcult and expensive. The transverse
spacing of twin I-girders and the side wing width of their slabs discourage the use of full-depth
precast deck panels, and two alternative solutions are typically adopted: in-place casting on
left-in-place forming systems, and in-place segmental casting with a self-positioning forming
carriage. In-place casting on conventional forms is such a slow and labour intensive process that
its use is limited to the shortest bridges.

4.6.2.1 Left-in-place forming systems


The costs and construction time related to slab casting on ground falsework are acceptable only in
particular cases. In most cases, the concrete slab is cast segmentally with a forming carriage

278

Downloaded by [] on [25/09/17]. Copyright ICE Publishing, all rights reserved.


Steel and composite bridges

(Rosignoli, 2013) or on left-in-place forms. Materials suitable for use as left-in-place forms are
steel, precast concrete, precast concrete acting compositely with steel beams or electro-welded
rebar lattices that are eventually embedded into the overlying in-place concrete, galvanised
proled steel sheeting (alone or welded to steel beams embedded into the in-place concrete), and
reinforced plastic or bre-cement sheeting or similar.

Some of these materials are of a fragile nature when used as left-in-place forms, and particular care
must be exerted to prevent accidents. Galvanised proled steel sheeting has been frequently used in
Japan to obtain concrete-lled I-beam grid decks (Kurita et al., 1998; Matsui et al., 1997). In spite
of the weight saving and the possibility of assembling semi-prefabricated formingreinforcement
panels on the steel girder prior to launching, these forms are rarely used on transverse spans longer
than 34 m, and precast concrete panels are often less expensive on these spans.

Composite slabs including a continuous 810 mm steel plate and a 70110 mm in-place concrete
overlay have been used in several suspended bridges and movable bridges because of their low
weight, which ranges between 2.5 and 3.5 kN/m2. The main limitation of the composite slabs is
related to the exibility of the steel plate during concrete pouring, which limits the transverse span
achievable to 1.52.5 m, and therefore requires multiple substringers. The steel plate is also exces-
sively thick for the service requirements of bending and shear (Bernard-Gely and Calgaro, 1994).

The left-in-place precast concrete forms are reinforced with full-width transverse electro-welded
rebar lattices or punched I-beams supported on the main girders. The transverse beams are
designed for self-weight and the load of uid concrete and personnel, and suspend precast concrete
panels for the side wings and the central slab strip between the main girders. The bottom chord/
ange of the transverse beams is embedded into the precast concrete panel, and the top chord/
ange is eventually embedded into the in-place concrete overlay. The vertical side bulkheads are
formed conventionally or with vertical precast concrete panels.

The precast concrete panels may or may not be considered as structurally participating with the
overlying in-place concrete under the action of loads imposed on the two-layered slab after casting.
No matter what the intent of the design is, the two-layered slab is highly sensitive to the time-
dependent properties of concrete.

Drying and thermal shrinkage causes cracks at the interface between precast panels and overlaying
in-place concrete. When the composite action between precast panels and in-place concrete is
relied upon, some design standards restrain the design of the composite two-layered slab with
stringent specications related to durability. Wheel loading affects the precast panels directly, and
the durability of reinforcement is a prime concern. This is particularly true when the deck has
spaced I-girders without substringers that create additional support lines for the panels.

Concrete cracking at the interface may affect durability, since humidity may pass through full-
depth cracks in the in-place concrete overlay and gather at the interface. If the top and bottom
concrete cover to panel reinforcement comply with the design standards, the panels are soon too
heavy for handling without cracking. The top rebar layer in the panel may be hot-dip galvanised or
epoxy coated to minimise the top cover, but a minimum top cover is still necessary for robustness
under casting loads (Rosignoli, 1999). The thickness resulting from a technological top cover, a
rebar grid designed for longitudinal and transverse positive bending in the sandwich slab, and a
code-complying bottom cover is often excessive as well.

279

Downloaded by [] on [25/09/17]. Copyright ICE Publishing, all rights reserved.


Bridge Launching

Thick panels also result in thin in-place topping, which may be affected by fatigue at the joints
between panels. Structural continuity between precast panels may be achieved with lapping
reinforcement protruding from the units, post-tensioning, embedded steel plates to be eld-welded
after panel assembly, or high-strength bolts. All these solutions may be affected by fatigue under
repeated loading, and are expensive and labour intensive. For all of these reasons, precast panels
are often considered as structurally non-participating.

When the panels are structurally non-participating, account is taken of the effects of differential
shrinkage and composite action that may adversely affect the bridge. Some design standards
specify that the requirements for concrete cover and crack control applicable to monolithic
concrete slabs must be satised in the in-place concrete layer, ignoring the presence of the precast
panels. The clear distance between panels and reinforcement embedded in the overlay should also
exceed the maximum nominal size of the aggregate.

The weakest point of non-participating panels is their weight. This additional load does not
produce any advantage in terms of exural capacity of the composite cross-section. Since
haunched panels are expensive, the two-layered slab typically has constant thickness in the central
bay between the main girders, and this further increases the weight of the decking system. A
composite bridge with 60 m spans, 7 m-spaced twin I-girders and a 13 m deck breadth may require
a 250 mm in-place layer, 50 mm non-participating precast panels and steelwork for 2.0 kN/m2.
With the same exural capacity, the cross-section weighs 9.5 kN/m2 in the presence of precast
panels and 8.3 kN/m2 without. If the superimposed dead load and the live load are, say,
7.5 kN/m2, the design load of the bridge increases from 15.8 kN/m2 to 17.0 kN/m2, and the weight
of the steelwork increases by 8%.

Precast panels are also expensive, handling and shipping involve additional costs, placement onto
the steel girder requires access under the bridge and is also expensive, and providing a suitable
seal between the steelwork and the panels to minimise grout leakage and the risk of corrosion
adds labour costs. When the piers are short and the area under the bridge is accessible, the
panels are positioned with ground cranes. Tall piers or ground obstructions require the use of
motorised trolleys that load multiple panels, roll along the steel girder and place the panels into
position.

Precast panels are often assembled directionally, from one abutment toward the other, to minimise
interference and facilitate access to the working area. Slab reinforcement may be assembled in the
same direction for an earlier start, and continuous slab casting in the same direction would be
the most intuitive and simplest solution. This casting method causes permanent tensile stress in
the negative bending regions when span segments are cast.

To limit cracking, segmental slab casting is specied according to the discontinuous sequence in
Figure 4.29. Span segments are cast in multiple bays, followed by pier segments. These casting
sequences reduce the self-weight tensile stress in the concrete. After the onset of composite action,
slab cracking is only related to the superimposed dead load, the live load and time-dependent
effects.

Step-by-step time-dependent analysis is necessary for the entire casting sequence. When a new
segment is cast in a span, cracking tends to occur at the end of the hardened slab segment in the
previous span. Camber analysis is also complex because of differential concrete creep between

280

Downloaded by [] on [25/09/17]. Copyright ICE Publishing, all rights reserved.


Steel and composite bridges

Figure 4.29 Segmental casting sequence for the concrete slab

1 2

3 4

5 6

the slab segments. The camber is analysed with modular ratios appropriate for the stage of
construction and the type of loading. An effective width is used for the concrete slab compatible
with the expected shear-lag effects.

4.6.2.2 Forming carriages


Medium-length bridges often allow for the investment of a self-positioning forming carriage for
the in-place casting of slab segments (Rosignoli, 2013). In continuous bridges with 6090 m spans,
the length of the slab segments is typically one-third of the span, with construction joints at the
span sixths and at the midspan. Discontinuous casting sequences are used where two span seg-
ments are cast in 45 spans followed by the pier segments, to reduce the permanent tensile stress
in the negative bending regions (Figure 4.30). The segments are cast monolithically in one pour.

A forming carriage consists of a casting cell suspended from an overhead frame that rolls along the
steel girder. In the rst-generation machines, the overhead frame is supported on four hydraulic
legs during repositioning and on multiple adjustable spikes during the casting of the slab segment.
Launch bogies and spikes are supported on rectangular rails welded to the anges of the steel gir-
der among lines of headed stud connectors (Figure 4.31). Rails and spikes are embedded into the
slab, and the spikes are extracted prior to repositioning of the carriage.

In some new-generation carriages, hybrid connections are used between the steel girder and the
concrete slab. Headed stud connectors are combined with workshop-welded steel shapes that
support the forming carriage during operations and repositioning (Figure 4.6). Removable
cross-fall shims are applied to the connectors to support xed launch bogies that set the rolling
plane for the carriage. The cross-fall shims form a recess in the slab surface that is lled with
non-shrink mortar after removal of the shim, to create the concrete cover for the connector.
The forming carriages for xed launch bogies are equipped with full-length support frames that
roll over the launch bogies during repositioning.

281

Downloaded by [] on [25/09/17]. Copyright ICE Publishing, all rights reserved.


Bridge Launching

Figure 4.30 Discontinuous slab casting

Figure 4.31 Hydraulic forming carriage for 22 m-long, 14.3 m-wide slab segments

282

Downloaded by [] on [25/09/17]. Copyright ICE Publishing, all rights reserved.


Steel and composite bridges

Figure 4.32 Internal view of the casting cell

The overhead frame includes longitudinal trusses over the steel girder, transverse trusses over end
bogies and interior support legs, and lateral and diagonal bracing. Square tubes and bolted eld
splices are used to stiffen the frame. The frame may support a covering and lateral enclosures to
protect the casting cell in cold weather. The casting cell includes a central form table and outer
forms for the side wings. Hanger bars suspend the central form table and wing forms from the
overhead frame (Figure 4.32). Multiple support points lighten the forms and the carriage, but the
large number of holes in the concrete slab increases the nishing costs. The hangers are applied
during cage assembly, and typically prevent cage prefabrication.

During repositioning, the central form table rolls on bogies applied to the top chord of the cross-
frames of the steel girder. Segment geometry is repetitive, and the casting cell does not require
major geometry adjustment. When the concrete strength has been reached, the hangers are
released and lifted, and the form table is lowered on the launch rolls for extraction. Forward
extraction is used for casting the next span segments; backward extraction and repositioning for
several spans is needed for casting the pier closures. The forms for the side wings are suspended
from the overhead frame, and rotated to vertical prior to repositioning of the carriage.

The overhead frame is assembled on the ground and lifted onto the steel girder in one operation.
The wing forms are applied with a symmetrical sequence, to minimise load imbalance, and the
carriage is moved to the rst casting location. Driving the carriage on multiple spans is a fast
operation when the frame has hydraulic legs. If the carriage has support frames for repositioning
on xed launch bogies, cross-fall shims for 34 spans are necessary for the typical four-span
casting sequence of pier segments. The modules of the central form table are lifted on the launch
rolls and connected to each other, and the form table is moved under the carriage.

283

Downloaded by [] on [25/09/17]. Copyright ICE Publishing, all rights reserved.


Bridge Launching

Figure 4.33 A 30 m forming carriage for 80 m spans. (Photo: Doka)

The carriage is as light as possible for cost reasons and to reduce the load applied to the steel
girder. The carriage is stable longitudinally and does not require counterweights, but it is rarely
anchored to the steel girder, and overturning may occur in the transverse plane. The risk of
overturning increases as the spacing of the steel girders diminishes and the width of the side wings
increases. Overturning is controlled with lling sequences that include casting the central strip rst,
and lling the side wings symmetrically afterward. A U-girder is exible in torsion and distortion
prior to achieving composite hollow-section behaviour, and symmetrical casting sequences also
control twist and instability.

The segments are cast in one phase without cold joints. Concrete is typically pumped from the
ground. On-deck concrete delivery may be necessary when the bridge has tall piers on inaccessible
terrain (Figure 4.33). Temporary steel bridges supported on the steel girder are necessary at
the solutions of continuity over the piers, to stabilise the steel girder and provide access for the
concrete mixers.

The concrete slab is cast directionally from abutment to abutment. The casting sequence involves
shuttling the forming carriage and the central form table back and forth along the deck. When the
carriage is equipped with hydraulic legs, motorised wheels drive narrow launch bogies along the
launch rails without interfering with the stud connectors. When cross-fall shims and xed launch
bogies are used, the support systems of the carriage are above the slab, and the geometry is less
constrained.

A 2030 m forming carriage with a power-pack unit and hydraulic systems is not inexpensive. The
self-weight may be 2.02.5 kN per square metre of formed surface. The costs of shipping, site
assembly and decommissioning also depend on the weight of the machine. Although the contrac-
tors depreciation strategy ultimately denes the cost share charged to the project, bridges shorter
than 150200 m are rarely compatible with such investments.

284

Downloaded by [] on [25/09/17]. Copyright ICE Publishing, all rights reserved.


Steel and composite bridges

Figure 4.34 A 25 m forming carriage for span segments and 15 m carriage for pier segments.
(Photo: Doka)

The productivity is one slab segment per week after the learning curve has passed. The incremental
launching of the steel girder followed by in-place slab casting with a forming carriage is one of the
fastest construction methods for medium-span bridges. When the bridge is not very long, the
reinforcement cage can be assembled behind the abutment and incrementally launched into
position to reduce the cycle time of the forming carriage (Bernard, 1997). The support systems for
the carriage are applied to the steel girder on launch completion of the steel cage.

Forward repositioning of the carriage and the form table takes a few hours, and the geometry
adjustment is very simple. Shuttling the carriage and the form table back and forth along the
bridge takes more time and disrupts the casting cycle. When the length of the bridge justies the
investment, a second carriage may be used for shorter pier segments (Figure 4.34). Fixed launch
bogies on cross-fall shims are the rst-choice solution when two carriages are used on the deck
(Rosignoli, 2013).

4.6.2.3 Control of cracking


Without preventive measures, the concrete slab of a composite bridge tends to develop cracking
over time. Early cracks before hardening are due to plastic settlement and plastic shrinkage.
These two phenomena may cause cracking that is of signicance, but preventive measures can
control them. Continuous slab casting from abutment to abutment may cause cracking during
construction, and discontinuous casting sequences may be specied. However, after hardening
there exist at least ve additional causes of cracking: external loads, creep, drying shrinkage,
external temperature variations and thermal shrinkage (Cremer et al., 1997; Navarro and
Lebet, 2001). The design standards cover most of these phenomena. The thermal shrinkage of
concrete is typically disregarded, although it may be a prime cause of early cracking of cast-in-
place slabs (Ducret and Lebet, 1997, 1999).

285

Downloaded by [] on [25/09/17]. Copyright ICE Publishing, all rights reserved.


Bridge Launching

During cement hydration, the temperature in the concrete slab increases by 15308C during the
rst 1224 h. This heating phase is followed by a cooling period of 150180 h. Cement hydration
governs strength development (Collepardi, 1980), and the modulus of elasticity of concrete is
higher in the cooling period than in the heating phase.

If the concrete slab and the steel girder act compositely from the moment that the concrete sets,
then this composite action prevents the thermal expansion of the concrete during the heating phase
and its contraction during the cooling period. The restraint that the steel girder applies to the
concrete slab can be modelled by assuming constant but different values for the elastic modulus
of the concrete during the heating and cooling periods. Restrained thermal expansion during heat-
ing generates compressive strains in the concrete that are recovered during cooling. The difference
in the modulus of elasticity of the concrete in these two phases generates permanent tensile stresses
that may be high when compared with the tensile strength of young concrete.

The restraint exerted by the steel girder can be represented by the ratio b = As /Ac of the cross-
sectional areas of the steel girder, As , and of the concrete slab, Ac (Ducret and Lebet, 1997,
1999). The current trend to reduce labour costs at the expense of increasing the quantity of steel,
and with it the restraint that the steel girder exerts on the concrete slab, increases the practical
relevance of the effects of thermal shrinkage.

Typical values of b for recent twin I-girder decks increase from 0.04 to 0.12 when the span length
increases from 30 to 80 m. Laboratory tests and numerical analysis showed that the residual tensile
stress sr in a concrete slab varies in the range 0.50.8 MPa for sections with b = 0.05, 1.21.5 MPa
for sections with b = 0.08 and 1.62.0 MPa for sections with b = 0.12 (Ducret and Lebet, 1997,
1999). These stress levels are signicant when compared with the 2.02.5 MPa tensile strength
of young concrete, and can lead to early full-depth cracking. They also reduce the available tensile
strength of concrete, and an effective tensile strength should be used when determining the stiff-
ness of the composite section.

The following relationship has been proposed for the residual tensile stress in the concrete slab
(Ducret and Lebet, 1997):

b2 Es2 Ec,h Ec,c
sr = a DT (4.31)
bEs + Ec,h bEs + Ec,c

where a is the coefcient of thermal expansion of concrete, Ec,h and Ec,c are the mean values for
the modulus of elasticity of concrete during the heating phase and cooling period, respectively, Es
is the elastic modulus of the steel girder, and DT is the maximum difference between the ambient
temperature and the concrete temperature during cement hydration.

A qualitative evaluation of the inuence of b on the effects of thermal shrinkage led to the
following observations. Hydration effects have limited inuence on early cracking for b 0.05.
For 0.05 , b 0.08, thermal shrinkage reduces the effective tensile strength of concrete and leads
to a limited risk of early cracking. For 0.08 , b 0.12, thermal shrinkage reduces the effective
tensile strength more markedly and early cracking is probable, and actions for reducing the
residual tensile stress should be considered. For b . 0.12, thermal shrinkage signicantly reduces
the effective tensile strength and a high risk of early cracking arises, and actions for reducing the
residual tensile stress should be adopted.

286

Downloaded by [] on [25/09/17]. Copyright ICE Publishing, all rights reserved.


Steel and composite bridges

Table 4.1 Tensile stress components (MPa) in the concrete slab at the interior supports

30 m span 80 m span
b = 0.04 b = 0.12

Thermal shrinkage 0.8 2.0


Casting sequence (continuous method) 1.8 2.7
Road surfacing 0.8 1.3
Traffic 0.3 0.1
Drying shrinkage 0.8 1.4

Possible corrective actions are all addressed at controlling DT. They involve reductions in the
cement content and the water-to-cement ratio (Berthellemy et al., 1997a), the use of low-heat
cement, cooling of fresh concrete or, during hydration, by cooling pipes, and slight heating of the
steel girder. Laboratory tests showed that the use of low-heat cement and concrete cooling can
limit the residual tensile stress in the concrete slab to less than 1.0 MPa even in the case of a high
restraint level (b = 0.12).

For technical and cost reasons, most corrective actions are impractical, and the typical solution is
the careful design of passive reinforcement. In this case, full-depth cracks will open, and it will
often be impossible to limit their width to below 0.1 mm. Detailing of the bar diameter and
distribution often permits limiting the crack width to 0.150.2 mm, but the crack width may
increase with time because of progressive bond deterioration at the crack boundaries due to fatigue
effects of live loading and temperature variations, with repercussions on structure durability.

Table 4.1 summarises, according to their chronological occurrence, the concrete slab tensile
stresses in the negative bending regions of composite bridges (Ducret and Lebet, 1999). Results
are given for concrete slabs cast with a continuous pouring sequence from one abutment to
another. The effects of the external temperature are not included. The tensile stress due to trafc
has been evaluated for a single 250 kN truck, and the effects of shrinkage have been measured on
actual bridges. The table shows that, because of the low early tensile strength of concrete, the
effects of thermal shrinkage and the casting sequence can initiate cracking in the concrete slab
of 30 m-span bridges, and can cause signicant cracking in 80 m-span bridges.

4.6.2.4 Longitudinal prestressing


In the presence of diffused full-depth cracks, the primary protection of the concrete slab is the use
of a waterproof membrane of good quality, rmly anchored to a sound concrete cover with low
permeability, and perfectly sealed at the joints and the edge barriers. A waterproof membrane
controls the two main causes of most chemical and physical processes affecting the durability
of structures: transport within pores and cracks, and water. Strength deterioration due to
freezethaw cycles, the effects of de-icing salts, the penetration of chemically aggressive agents, the
alkalisilica reaction, concrete carbonation and reinforcement corrosion are all related to these
two causes.

The stability of the protective action of a waterproof membrane over time is a basic concern. Despite
all precautions, the use of de-icing salts will sooner or later result in the penetration of chloride ions
into the concrete. As the slab is cracked, the rate of penetration is high, and de-passivation occurs

287

Downloaded by [] on [25/09/17]. Copyright ICE Publishing, all rights reserved.


Bridge Launching

quickly. In these circumstances, structure durability is controlled by corrosion development.


Service lives of 2040 years have been mentioned in the literatue, but this should be conrmed
by additional research.

Applying prestressing is a practical way of improving the durability of the concrete slab. Slab pre-
stressing can reduce or prevent concrete cracking, thus increasing the stiffness of the slab and
improving the protection of reinforcing steel from corrosion. Slab prestressing can be applied in
the longitudinal direction, in the transverse direction, and simultaneously in both directions.
The prestressing concepts follow two philosophies derived from the PC bridges.

g Total prestressing aims to avoid cracking under the combined effects of the self-weight, the
superimposed dead load, the live load, differences in temperature and time-dependent
effects such as losses of prestress, creep and shrinkage. Small tensile stresses may be allowed
at the edges of the slab, and the slab is analysed as a homogeneous non-cracked section.
g Partial prestressing is designed to resist permanent and time-dependent effects with non-cracked
sections, while some cracking is accepted under live loading. Limitations are placed on the
crack width by specifying the rebar distribution and limit values for the SLS tensile stress in the
reinforcement steel in relation to the concrete cover and the aggressiveness of the environment.

Jacking of the interior supports of the continuous beam has been used frequently to apply
longitudinal compression to the concrete slab. The supports are raised prior to slab casting, and
lowered after the concrete has cured and a stable composite action has been achieved. Positive
bending is thus introduced in the composite deck, and the concrete slab is compressed. The
effectiveness of countering shrinkage is only partial, as the benets are balanced by creep. The
time-dependent stress redistribution within the composite deck depends on the relative stiffness:
the stiffer the steel girder, the higher the loss of slab compression over time. Therefore, this system
of slab prestressing is rarely used for bridges with large steel sections.

Other methods involve temporary prestressing of the steel girder. The stability of the steel girder
during prestressing is a prime concern. When the slab is compressed by releasing temporary
prestress in the steel girder, the compressive stress in the concrete should not exceed 5060% of
the strength at transfer. If the composite deck is prestressed by an external system or by tendons
not directly bonded to the concrete, the prestressing forces are calculated accounting for the
deformation of the whole structure.

Longitudinal prestressing with internal tendons in the concrete slab is seldom employed because it
adds considerably to the bridge cost. While transverse slab prestressing may result in some weight
saving, longitudinal prestressing produces minimal structural benets, and is introduced mainly
for durability purposes. Prestressing tendons are not required for the ultimate negative exural
capacity of the composite cross-section, as the top anges of the steel girder and slab reinforcement
designed to limit the crack width are typically sufcient to ensure adequate capacity.

Further problems derive from cracking at the anchorages of post-tensioning cables, and from the
difculty of controlling the losses of prestress due to the composite action. Part of the prestressing
is lost in the steel girder at tensioning, and the loss of prestress grows with time due to creep of
concrete. The time-dependent redistribution of the prestress depends on the relative stiffness,
and since longitudinal prestressing is mainly used for long spans with large steel sections, the
limitations are evident.

288

Downloaded by [] on [25/09/17]. Copyright ICE Publishing, all rights reserved.


Steel and composite bridges

Although composite bridges have become standard practice in several countries and many design
standards are available, few include provisions for longitudinally prestressed concrete slabs. In
most cases, engineers must resort to design standards for PC and structural steel. Prestressing and
time-dependent effects such as creep and relaxation do not affect the ultimate exural capacity of
the composite cross-section under negative bending. The plastic rotation capacity decreases
abruptly when the web slenderness increases, and the effects of local buckling in the steel girder
must therefore be assessed. The fatigue strength must also be evaluated, and because the stress
range is a controlling factor, the design must address the live load effects. This is more important
in composite decks made with high-grade steel, as no increase in the fatigue strength should be
expected at welded details and connections.

For all of these reasons, the current trend is to avoid longitudinal slab prestressing unless depth
restrictions control, slab cracking must be prevented, or structural continuity must be provided for
a concrete slab composed of full-depth precast panels with non-reinforced grouted joints.
Incremental launching of the concrete slab over the steel girder solves most of these limitations
and makes longitudinal prestressing more affordable.

4.6.2.5 Transverse prestressing


Transverse prestressing is introduced to control cracking and creep deections in the concrete slab
under self-weight and superimposed dead loads. Either full or partial prestressing may be used.
Post-tensioning tendons are used in the monolithic slabs cast in-place with forming carriages, and
pre-tensioned strand is used at the mid-plane of precast deck panels or as their edge reinforcement.

Transverse post-tensioning is expensive, but has been used occasionally for wide twin I-girder
bridges. It may become cost-effective when the deck width exceeds 1518 m, as the transverse
spacing of the I-girders may be increased to 810 m. Prestress losses and the cost of skilled labour
and anchorages rarely make transverse post-tensioning cost-effective in narrower twin I-girder
bridges.

Transverse post-tensioning is used more frequently in the concrete slab of composite box girders.
The high exural capacity and torsional stiffness of a composite box girder allow the use of narrow
U-girders (Berthellemy et al., 1997b), which are light and easy to handle. The side wings of the
concrete slab are wider than in twin I-girder decks, and transverse slab prestressing saves weight
and controls the exibility of the cross-section.

Compared with transverse prestressing of PC bridges, the tendons are less powerful (mono-strands
with prestressing forces of 150160 kN) and spaced less frequently (0.20.6 m) in relation to the
distribution of reinforcement and the intensity of transverse bending. Paired mono-strands at
0.8 m spacing may reduce the reinforcement ratio from 250 kg/m3 to 180190 kg/m3 (Berthellemy
et al., 1997b). The concrete slab is thin, and the tendons are often designed with a straight layout.
Stressing anchorages are placed alternately on either side of the slab. Transverse post-tensioning
of wide slabs can be obtained with four-strand tendons placed in transverse ribs, with additional
stiffness advantages.

4.6.3 Incremental slab launching over the steel girder


In many instances, the cost of full-depth precast deck panels, two-layer slabs on precast con-
crete panels and monolithic slabs cast in-place with forming carriages warrants consideration of
alternatives. An alternative construction method for the concrete slab of a composite bridge

289

Downloaded by [] on [25/09/17]. Copyright ICE Publishing, all rights reserved.


Bridge Launching

Figure 4.35 Segmental slab casting and incremental launching

Mobile covering Casting cell

Rebar jig

with a constant and large plan radius involves segmental casting of a full-length continuous
slab behind the abutment combined with incremental launching of the slab over the steel girder
(Figure 4.35).

The casting cycle includes casting a 2025 m slab segment on a casting bed, jacking the entire
slab forward to clear the casting bed, and splicing reinforcement and prestressing ducts of the new
segment with the rebar and ducts protruding from the rear end of the launched slab. On launch
completion, the slab is connected to the steel girders by mechanical connectors, welding, lling
of continuity pockets over the girders and the like. The advantages are repeated use of the same
formwork, concrete processing in a xed and sheltered facility, and a structurally continuous
monolithic slab.

The casting bed may be supported on the ground behind the abutment or may be suspended from
the steel girder at any location. A xed casting bed is much less expensive than a forming carriage,
the slab is cast with a continuous process, and the logistics are substantially simplied. Slab
launching reduces the risks for workers, avoids interference with the area beneath the bridge,
reduces shrinkage and creep cracking, facilitates the application of longitudinal post-tensioning
in the negative bending regions, simplies structural analysis and improves durability.

4.6.3.1 Technology
In the rst applications of this construction method, the slab was launched directly over the anges
of the steel girder. When the slab reached its nal position, shear connectors were welded to the
anges within stud pockets that were eventually lled with non-shrink concrete or grout. This
solution is affected by some limitations. The eld welding of shear connectors within small open-
ings is expensive, lling stud pockets distributed throughout the length of the bridge is also
expensive, localised connection forces tend to cause cracking with time, and the lateral support

290

Downloaded by [] on [25/09/17]. Copyright ICE Publishing, all rights reserved.


Steel and composite bridges

Figure 4.36 Slab launch on the flange plates and continuous connection

of the compression ange is localised at few points. Guiding the slab during launching is also
difcult with this type of girderslab connection.

During launching, the concrete slab is supported on launch shoes that slide along the edges of the
anges. High sliding friction at inaccessible support points generates anomalous stresses and
increases the cost of the thrust devices. Irregular support lines for the concrete slab due to the
cambered prole of the girder cause stress concentration and risks of cracking. Transverse bending
and twist in the top anges may require thicker plates and vertical web stiffeners to protect the
weld of the web to the ange, and splice plates and bolted joints in the top ange prevent slab
launching.

After some trials, continuous launch openings lled with non-shrink concrete at the end of launch
were introduced into the slab (Figure 4.36). Continuous openings avoid the concentration of
connection stresses and improve the lateral stability of the compression ange, but do not solve
the other weak points of the original scheme, and actually create new problems. Transverse
reinforcement connecting the slab strips must ensure stability of the side wings. The vertical lever
arm is small, the transverse forces are high, and buckling of the compression bars may cause over-
turning of the side wing. Relative displacements between the slab strips may also cause buckling of
the compression bars, and the slab strips must therefore be braced.

Solutions based on vertical steel plates embedded into the concrete slab and emerging under the
soft were also tested. On launch completion, the plates are welded to the ange edges to achieve
longitudinal continuity. In the embedded portion, the plates support headed stud connectors or
are punched with round holes to receive concrete dowels (Kraus and Wurzer, 1997). This solution
avoids solutions of continuity in the slab, improves structural durability, distributes the

291

Downloaded by [] on [25/09/17]. Copyright ICE Publishing, all rights reserved.


Bridge Launching

Figure 4.37 Indirect slab launching

Cross-beam

Launch rail

connection forces, and facilitates future slab replacement by pulling the slab backward after weld
grinding. However, this solution requires stringent dimensional tolerances in the concrete slab and
the steel girder, and extensive overhead welding from mobile platforms, and the launch surfaces
are not accessible. Lateral llet welds on one ange edge only prevent the transfer of transverse
bending between the slab and the anges, and may cause fatigue (Akesson et al., 1997).

The weak points of the solutions examined so far may be avoided by launching the slab on longi-
tudinal rails supported on the anges (Figure 4.37). The launch rails are welded to the anges with
interposition of bearing plates (Figure 4.38). Short cross-beams connect the slab strips through the
launch openings, and are supported on launch shoes that control friction (Rosignoli, 1998, 1999).
During launching, the weight of the slab is transferred to the steel girder through cross-beams,
launch shoes, launch rails and bearing plates.

The shear stiffness of the cross-beams facilitates transfer of the slab support reactions. Their axial
and exural stiffness, combined with the tensile capacity of transverse top reinforcement, controls
the stability of the cantilever slab strips and limits their rotation. In-plane exural stiffness and no
differential in the launch frictional resistance avoid the need for temporary bracing between the
slab strips. The bearing plates of the launch rails compensate for ange irregularities and localise
the loads applied to the steel girder, which may be stiffened if necessary; the launch shoes control
launch friction, and the sliding surfaces are immediately accessible.

On launch completion, the launch openings are completed with reinforcement cages, and non-
shrink concrete is poured to connect the slab to the girder without secondary stresses. The launch

292

Downloaded by [] on [25/09/17]. Copyright ICE Publishing, all rights reserved.


Steel and composite bridges

Figure 4.38 Launch opening

Cross-beam

Launch shoe
Top rebar
Launch rail
Stud connectors

Bearing plate

rails may be left in-place or recovered. The launch shoes are recovered, and the cross-beams are
shimmed for the nal control of elevation. In long bridges, the launched slab may be divided into
sections separated by transverse joints. Jacking at the joints is used to align the slab along the steel
girder in case of large thermal or time-dependent deformations. The temporary joints are also
lled with concrete on completion of the slab adjustment.

4.6.3.2 Advantages over in-place casting


The launch of a concrete slab over a steel girder is subject to the same geometry requirements as
any other bridge launching application. Slab launching is particularly cost-effective when the steel
girder is also erected by incremental launching, as the same installation and logistics are used
for construction of both components of the structure. Although slab launching is compatible
with multi-girder systems and girdersubstringer systems, it is particularly cost-effective with twin
I-girders because of the minimum number of launch openings and the optimised varying-thickness
design of the monolithic slab.

Launching the slab offers several advantages over in-place casting. The rst advantage is avoiding
the need for a forming carriage. A forming carriage is not inexpensive, and short bridges often do
not permit such an investment (Rosignoli, 2013). The investment for slab launching is limited to a
xed casting bed, a system of launch rails, and a pair of hydraulic cylinders that apply the thrust
force. Most of the deck is built on the ground, in a sheltered facility, in any weather conditions,
and under a tower crane. The presence of workers and materials on the deck is minimised, and
there is no need to create accesses and working areas on the deck or on the ground.

Two-layer slabs on precast concrete panels also avoid the use of a forming carriage, but the precast
panels are expensive and increase the weight of the steelwork. The launched slabs are monolithic
and fully participating, the exural efciency of the composite section increases, and the weight of
the steelwork decreases. Even neglecting these cost savings, avoiding the cost of precast panels,
shipping and handling may cover the investment for slab launching even in short bridges. A mono-
lithic slab is more stable and durable than a two-layer slab, concrete cover may be thicker, and the
slab soft can be easily inspected at segment extraction from the casting cell and over time.

293

Downloaded by [] on [25/09/17]. Copyright ICE Publishing, all rights reserved.


Bridge Launching

The aesthetic advantages are no less signicant. The launched slabs have a regular and ordered
aspect, and launching permits the adoption of haunched shapes and rounded corners with
minimum costs. Sandblasting and high-pressure washing are easier, and deeper surface nishing
such as board-marking and engraved textures is inexpensive because of the small form dimensions.

The reinforcement cage can be entirely prefabricated or handled in grids to improve labour
rotation. Concrete pouring with buckets is less expensive than pumping, the area beneath the
bridge is unaffected, safer work conditions and repetitive operations enhance the productivity
of personnel, and the slab can be cast under a tower crane for easier handling of materials.

Launched slabs undergo most dimensional variations due to drying and thermal shrinkage before
being connected to the steel girder. The use of non-shrink concrete for lling the launch openings
reduces the residual stresses in these slab areas. The expansion action of concrete, in contrast to the
slab strips, generates a diffused compression that prevents shrinkage contraction from causing
cracking. Creep stresses are low as well, due to the absence of longitudinal long-term stresses in
the slab and the dimensional stability resulting from the long curing time at joining.

Longitudinal prestressing improves durability and makes the slab contribution to the exural
capacity of the composite section constant along the deck, which simplies structural analysis
and makes its results more reliable. Longitudinal prestressing involves complex operations in a
cast-in-place slab, while its application is easy and inexpensive during the construction of a
launched slab. The loss of prestress in the steel girder is also smaller because of the long curing
time at joining.

The concrete slab is 34 times heavier than the steel girder. During slab launching, the load
eccentricity and load concentration may both be risky, and most of these risks are not readily
quantiable. Slab launching should always be carried out in close collaboration with the bridge
designer, and unexpected events should be recorded, investigated and shared between all parties
involved.

4.6.3.3 Stress analysis


Factors to be considered in the design are related to the slab launch process and to the effects of
the launch forces on the concrete slab and the steel girder. The jacking force is applied parallel to
the plane of the slab, and must overcome the frictional resistance of the launch rails. For a body
moving at low speed, the frictional resistance is discontinuous and is applied through a series of
sticks and slips. The stickslip effects depend on the weight of the slab and the steel girder, the
length of the launch strokes and the exibility of the launch shoes. Guiding the slab along the
launch path may also require the application of transverse forces.

The simple span shown in Figure 4.39 is subject to a uniformly distributed friction load of
intensity r. The span has a cross-section with the centre of gravity at mid-height. If the beam is
restrained longitudinally at point C, the eccentricity is small, and the only force acting at a distance
x is an axial force P = rx. The continuous beam shown in Figure 4.40 is restrained at the level of
the top ange (point F), and the internal forces may be determined as follows.

First, the continuous beam is restrained at mid-height, as shown in the middle of Figure 4.40,
where Ltot is the total length of the concrete slab supported on the steel girder. The only load
effect is the axial tension P = rLtot . Then, the continuous beam is subject to a terminal couple

294

Downloaded by [] on [25/09/17]. Copyright ICE Publishing, all rights reserved.


Steel and composite bridges

Figure 4.39 Simple span under uniform friction loading

x
r
C

Figure 4.40 Continuous beam under uniform friction loading

L L L

Ltot

F r

r
C

M = PH/2, as shown at the bottom of Figure 4.40. For n equal spans of constant exural stiffness,
the bending moment MJ over the support J is

sinh j g
MJ = M (1)n + j (4.32)
sinh ng

where cosh g = 2. If the jacking force is not symmetrical about the bridge centreline, the frictional
resistance of the two I-girders causes horizontal bending and warping. The launch jacks should
therefore be fed in parallel by a common hydraulic system.

REFERENCES
AASHTO (2003) Guide Specications for Horizontally Curved Highway Bridges. AASHTO,
Washington, DC, USA.
AASHTO (2012) LRFD Bridge Design Specications. AASHTO, Washington, DC, USA.

295

Downloaded by [] on [25/09/17]. Copyright ICE Publishing, all rights reserved.


Bridge Launching

Akesson B, Edlund B and Shen D (1997) Fatigue cracking in a steel railway bridge. Structural
Engineering International 7(2):118120.
Azizinamini A (2013) Design Guide for Bridges for Service Life. Transportation Research
Board, Washington, DC, USA. SHRP2.
Belluzzi O (1988) Scienza delle Costruzioni. Zanichelli Editore, Bologna, Italy.
Bernard J (1997) The Rocher de lEscalade Viaduct, France. Structural Engineering Inter-
national 7(1): 1920.
Bernard-Gely A and Calgaro JA (1994) Conception des Ponts. Presses de l Ecole Nationale des
Ponts et Chaussees, Paris, France.
Berthellemy J, Lacombe JM and Poineau D (1997a) Cracking control in the concrete slab of
the Nevers composite bridge. IABSE FIP-CEB International Conference, Innsbruck, Austria.
Berthellemy J, Lacombe JM and Poineau D (1997b) The Nevers Bridge: design of the steel
concrete composite box girder. IABSE FIP-CEB International Conference, Innsbruck, Austria.
Berthellemy J, Placidi M and Virlogeux M (1997c) Prefabrication and prestressing of concrete
slabs in composite bridges. IABSE FIP-CEB International Conference, Innsbruck, Austria.
BSI (British Standards Institution) (1982) BS 5400: Steel, concrete and composite bridges Part
3. Code of practice for design of steel bridges. BSI, Milton Keynes, UK.
Chemerinski OI, Seliverstov VA and Zhuravov LN (1996) Launching steel bridges in Russia.
Structural Engineering International 6(3): 183186.
CNR (Consiglio Nazionale delle Ricerche) (1988) CNR 10011: Strutture di acciaio. Istruzioni
per il calcolo, lesecuzione, il collaudo e la manutenzione. CNR, Rome, Italy.
Collepardi M (1980) Scienza e Tecnologia del Calcestruzzo. Hoepli Editore, Milan, Italy.
Contin A, Viviani E and Viviani M (2011) New Po River Bridge at Piacenza: the construction
process. IABSE Structural Engineering International 21(4): 433436.
Cremer JM, Dotreppe JC, Ernens M and Lothaire A (1997) Cracking and durability of concrete
slabs of composite bridges. IABSE FIP-CEB International Conference, Innsbruck, Austria.
CSA (Canadian Standard Association) (1994) CAN/CSA-S16.1-94: Limit states design of steel
structures. Canadian Standard Association, Rexdale, Ontario, Canada.
Ducret JM and Lebet JP (1997) Effects of concrete hydration on composite bridges. IABSE
FIP-CEB International Conference, Innsbruck, Austria.
Ducret JM and Lebet JP (1999) Behaviour of composite bridges during construction. Structural
Engineering International 9(3): 212218.
Dunker KF and Rabbat BG (1992) Performance of prestressed concrete highway bridges in the
United States the rst 40 years. PCI Journal 37(3): 4864.
Gale R (2011) Incremental launching of steel girders in British Columbia two case studies.
Structural Engineering International 21(4): 443449.
Geier R (2010) Design and construction of the Auenbach Bridge, Austria. Structural Engineer-
ing International 20(2): 153156.
Granath P (1998) Distribution of support reaction against a steel girder on a launching shoe.
Journal of Constructional Steel Research 47(3): 245270.
Granath P (2000) Serviceability limit state of I-shaped steel girders subjected to patch loading.
Journal of Constructional Steel Research 54(3): 387408.
Granath P, Thorsson A and Edlund B (2000) I-shaped steel girders subjected to bending
moment and travelling patch loading. Journal of Constructional Steel Research 54(3): 409421.
Hoeckman W (2001) Bridge over the River Loire in Orleans, France. Structural Engineering
International 11(2): 9498.
Horvath A, Dunai L and Nagy Z (2006) Dunaujvaros Danube Bridge: construction, design and
research. Structural Engineering International 16(1): 3135.

296

Downloaded by [] on [25/09/17]. Copyright ICE Publishing, all rights reserved.


Steel and composite bridges

Knight M and Fryer S (2011) The engineering and aesthetics of the Town Centre Link, London,
UK. Structural Engineering International 21(1): 5155.
Kraus D and Wurzer O (1997) Bearing capacity of concrete dowels. IABSE FIP-CEB Inter-
national Conference, Innsbruck, Austria.
Kurita A, Matsui S, Nakai H and Yoda T (1998) Trends in steelconcrete composite bridges in
Japan. Structural Engineering International 8(1): 3034.
Lockmann H and Marzahn G (2009) Spanning the Rhine River with a new cable-stayed bridge.
Structural Engineering International 19(3): 271276.
Malite M, Takeya T, Goncalves RM and de Sales JJ (2000) Monitoring of the Parana River
Bridge during launching. Structural Engineering International 3(3): 193196.
Matildi P and Matildi G (1990) Ponti Metallici, Esperienze Vissute. Cimolai, Pordenone, Italy.
Mato FM and Santos JP (2007) Arrojo Las Piedras Viaduct: the rst composite steelconcrete
high speed railway bridge in Spain. Structural Engineering International 17(4): 292297.
Matsui S, Mori K, Ohta K and Takagi M (1997) Design method and fatigue strength of large
span concrete-lled I-beam grid deck. IABSE FIP-CEB International Conference, Innsbruck,
Austria.
Monov B and Seliverstov Y (1997) Erection of composite bridges with precast deck slabs.
IABSE FIP-CEB International Conference, Innsbruck, Austria.
Navarro MG and Lebet JP (2001) Concrete cracking in composite bridges: tests, models and
design proposals. Structural Engineering International 11(3): 184190.
Navarro-Manso A (2013) Nuevo metodo de lanzamiento de puentes metalicos basado en doble
cajon colaborante: Simulacion numerica estructural y experimentacion aerodinamica. Univer-
sidad de Cantabria, Escuela Tecnica Superior de Ingenieros de Caminos, Canales y Puertos,
Spain.
NCHRP (National Cooperative Highway Research Program) (2008) Full-Depth Precast Con-
crete Bridge Deck Panel Systems. Transportation Research Board of the National Academies,
Washington, DC, USA. Report 584.
Petrangeli MP (1993) Progettazione e Costruzione di Ponti. Masson, Editoriale ESA, Milan, Italy.
Popov OA, Seliverstov VA (1998) Steel bridges on Ankara Peripheral Motorway. Structural
Engineering International 18(3): 205210.
Reis AJ (2001) Urban bridges: design, environmental and construction issues. Structural
Engineering International 11(3): 196201.
Romaro C and Romaro G (2000) Erection of arch and arch-frame bridges. Structural Engineer-
ing International 10(4): 259262.
Rosignoli M (1998) Il varo delle solette nei ponti compositi. LIndustria delle Costruzioni,
November.
Rosignoli M (1999) Estrusione delle solette nei ponti in acciaio e c.a. Le Strade, May.
Rosignoli M (2013) Bridge Construction Equipment. ICE, London, UK.
Rosignoli M (2014) Tappan Zee Hudson River Crossing: the new NY bridge project. Proceed-
ings of Madrid 2014 IABSE Symposium, Madrid, Spain.
RPX-95 (1996) Recomendaciones Para el Proyecto de Puentes Mixtas Para Carreteras. RPX-95,
Madrid, Spain.
Saul R (1996) Bridges with double composite action. Structural Engineering International 6(1):
3236.
Saul R (1997) Erection methods for long-span steel composite bridges. IABSE FIP-CEB Inter-
national Conference, Innsbruck, Austria.
Sayed-Ahmed EY (2005) Lateral torsionexure buckling of corrugated web steel girders. Pro-
ceedings of the Institution of Civil Engineers, Structures and Buildings 158: 5369.

297

Downloaded by [] on [25/09/17]. Copyright ICE Publishing, all rights reserved.


Bridge Launching

Stanton JF, Eberhard MO, Wacker JM and Hieber DG (2005) State-of-the-Art Report on Pre-
cast Concrete Systems for Rapid Construction of Bridges. Washington State Transportation
Commission, Federal Highway Administration, Washington, DC, USA. Contract T2695,
Task 53, Final Technical Report.
Timoshenko S and Gere J (1961) Theory of Elastic Stability. McGraw-Hill, New York, USA.
UNI (Ente Nazionale Italiano di Unicazione) (2005a) EN 1993-1-1:2005: Eurocode 3 Design
of steel structures Part 1.1: General rules and rules for buildings. UNI, Milan, Italy.
UNI (2005b) EN 1994-1-1:2004: Eurocode 4 Design of composite steel and concrete structures
Part 1-1: General rules and rules for buildings. UNI, Milan, Italy.
UNI (2006) EN 1994-2:2005: Eurocode 4 Design of composite steel and concrete structures
Part 2: General rules and rules for bridges. UNI, Milan, Italy.
Xanthakos PP (1994) Theory and Design of Bridges. Wiley, New York, USA.
Yura J (2001) Fundamentals of beam bracing. AISC Engineering Journal 38(1): 1126.
Yura J, Helwig T, Herman R and Zhou C (2008) Global lateral buckling of I-shaped girder
systems. Journal of Structural Engineering 134(9): 14871494.
Zhuravov L, Chemerinsky O and Seliverstov V (1996) Launching steel bridges in Russia. Struc-
tural Engineering International 6(3): 183186.

298

Downloaded by [] on [25/09/17]. Copyright ICE Publishing, all rights reserved.

You might also like