Sei sulla pagina 1di 120

Bridge Launching

ISBN 978-0-7277-5997-9

ICE Publishing: All rights reserved


http://dx.doi.org/10.1680/bl.59979.017

Chapter 2
Bridge launching

Incremental launching is a time-tested and extremely versatile technology. It has been applied to
pedestrian, highway and railway bridges. It has been applied to prestressed concrete (PC) solid
slabs, voided slabs, ribbed slabs with double-T section, single- and multi-cellular box girders, twin
box girders, and long-span Vierendeel girders. It has been applied to twin steel I-girders, multi-
girder systems, girdersubstringer systems, U-girders for composite box girders, steel girders and
trusses with orthotropic deck, and prestressed composite box girders with steel corrugated-plate
webs. It has been applied to single-span overpasses and to multispan bridges longer than 1 km;
to arch bridges, tied arches, cable-stayed bridges and aqueduct bridges; and even to a 46 m-wide
navigable channel bridge weighing 65 000 tonnes (metric tons) and carrying 80 000 tonnes of
water for transit of heavy barges.

The deck is launched as a continuous beam, but this is only the rst step of construction, and a
great number of alternatives for the nal structural conguration are possible. The deck may be
separated into shorter continuous beams or simple spans, and continuous beams launched from
the opposite abutments can be connected at the centre of the bridge to generate a longer beam. The
deck can be launched over an arch, or can be launched over temporary piers and suspended from
an arch, or suspended from one or more towers to create a cable-stayed bridge.

Incremental launching alternates deck launching with the construction of new segments. Bridge
launching is not necessarily incremental though. A great number of simple spans have been cast
or assembled behind the abutment and launched over the obstruction to overpass in one oper-
ation. These bridges include PC spans, steel spans, and composite spans and bowstring arches.
Launching may also be combined with jacking and skidding.

The extensive range of available techniques, the exibility and the possibility of combining
multiple techniques are the major advantages of this group of concepts. These construction
methods have an excellent reputation in relation to safety of operators and the public, with a mini-
mal number of accidents. Sometimes these construction methods are chosen for direct cost
savings; other times they are chosen to minimise the impact of construction on the public and
to optimise the mobility of people and goods.

2.1. Launch of single spans


Several single spans have been cast behind one abutment and launched into position in one
operation. The spans are prone to overturning during launching, and numerous solutions have
been used to stabilise the operation. When a temporary pier is to be placed at the centre of the
span, launching requires only a short launch nose. When the obstruction to overpass cannot be
disrupted, a longer launch nose may be used to control overturning, or a temporary mast may
be used on the opposite abutment to suspend the front end of the span with stay cables. When the

17

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


Bridge Launching

Figure 2.1 Braced PC launch nose

span is launched over a highway, a self-propelled modular transporter (SPMT) platform may also
be used to support the front end.

The PC span of the Reggiolo Bridge in Italy was launched over an electried railway with a tem-
porary pier and a short PC launch nose (Rosignoli, 2007). The span has a complex trapezoidal
geometry, is 26.0 m long, and the width varies from 17.7 to 23.8 m. During launching, the four-cell
box girder was supported under the inner webs, and the edge girders were unsupported.

The casting yard included three full-length foundation beams beneath the inner webs of the span.
Three braced PC launch noses were used to control negative bending and overturning (Figure 2.1).
Because of the varying width of the span, a guide rail was embedded in the deck soft to control
lateral drift. The span was launched on hydraulic launch bearings by means of a pair of rear thrust
cylinders anchored to the central foundation beam.

Launching a single span is simpler than incremental launching and requires the same type of
equipment. The span is jacked before launching to insert the launch pads and to determine the
location of the centre of gravity (Figure 2.2). The span may be launched with xed or movable
bearings. Fixed bearings on micropiles are used on settling soils, the span must have constant
depth, and bending varies due to the varying length of the front and rear cantilevers. Movable
bearings are integral with the span during launching; they require stiff, full-length runway beams,
but the span depth may vary, and less launch prestressing is needed in the deck.

Single spans with a long front cantilever may be built on both sides of the obstruction to overpass
and launched symmetrically into position to get a three-span continuous beam after midspan
closure and application of continuity prestressing. The length of the end spans is 6070% of the
main span in order to control overturning during launch. The deck has constant depth, and each

18

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


Bridge launching

Figure 2.2 Pre-launch weighing of support reactions with displacement sensors

deck section is launched on three xed supports: one at the centre of the casting area, one at the
abutment and one at the inner pier. No launch noses are necessary for launching.

Steel spans are lighter and more exible than PC spans, but the launch technology is the same.
Several tied arches have been erected behind the abutment and launched into position. The deck-
ing system usually includes steel edge girders connected by oorbeams and a concrete deck made
of full-depth precast panels and in-place closure pours over edge girders and oorbeams. The edge
girders act as tension ties for the two arch ribs and as runway beams during launching.

The 65.5 m tied arch launched over a DB railway in Ebealbach, Germany, includes two vertical
box ribs braced at the top. The span is skewed to the railway line, and the arch ribs are therefore
staggered longitudinally. Two xed launch bogies at the abutment were combined with two rear
mobile bogies rolling along runway beams; additional launch bogies were used at the three
temporary piers (Romaro and Romaro, 2000).

The 70.4 m tied arch launched over the Adige River in Italy includes inclined box ribs braced at
the top. Two launch bogies at the abutment were combined with two rear mobile bogies on
runway beams. After reaching the overhang allowed by a rear counterweight, a long launch nose
supported on the other abutment was connected to the edge girders for launch completion without
temporary piers (Romaro and Romaro, 2000).

When the span is launched over an existing highway, the front end of the span may be supported
by SPMTs. The span may also be constructed close to the bridge and moved into position with
an integrated SPMT platform. SPMTs are often used when replacing existing crossings, as the
same SPMT platform is used to remove the old span and to install the new span, with minimal

19

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


Bridge Launching

Figure 2.3 SPMT delivery of the Rialto Bridge in Venice. (Photo: Fagioli)

disruption of trafc. SPMT integrated platforms on barges may be used for marine operations
(Figure 2.3).

SPMTs are robust computer-controlled platform vehicles that are able to lift, transport and place
heavy and large loads. They are used for many ultra-heavy transportation operations in the
industrial eld. In the bridge industry, SPMTs are used for handling spans with weights ranging
from a few hundred to more than 10 000 tonnes. SPMTs do not prevent span twisting, and are
therefore used in applications where the potential effects of loss or of unequal or uneven support
are acceptable (Rosignoli, 2013).

A stayed mast on the opposite abutment may be used to support the front end of the span during
launching. The Pavilion Bridge in Spain is a three-dimensional steel structure combining an
exhibition building and a footbridge over the Ebro River (Perez Perez et al., 2011). The 138.6 m
launched span weighs 2100 tonnes and is asymmetric, curved in plan, and of varying width and
height. In the rst 44 m of launch the span was supported on two lines of mobile bearings sliding
along runway beams. When the front skid shoes reached the end of the runway beam, the front end
of the span was supported by two 8.5 MN stay cables anchored to the top of a 40 m mast. The
longitudinal component of the pull in the stay cables was balanced by front mast retaining cables
and rear retaining cables anchored to a concrete block (Figure 2.4).

2.2. Rotation
Rotation can occur in the vertical plane and in the horizontal plane. Vertical rotation has been
used in several arch bridges, where the ribs were slip-formed in a quasi-vertical alignment and

20

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


Bridge launching

Figure 2.4 Launch with stayed mast and retaining cables. (Photo: Dragados)

lowered symmetrically for midspan closure. Vertical rotation has also been used for the inclined
legs of steel portal frames, which were delivered on the approach spans and tilted and lowered into
position, or erected vertically and rotated into position, or barged horizontally under the bridge
and lifted into position by rotation.

Rotation in the horizontal plane is mostly applied to three-span bridges where the central span
is twice as long as the end spans. One-half of the bridge is built on either side of the obstruction
to overpass. After completion, the two structures are rotated about the inner piers to the nal
alignment, and are connected by a midspan closure pour and continuity prestressing. This method
is mainly used to span over highways, railways and navigable channels, and avoids balanced
cantilever construction with form travellers over the obstruction (Rosignoli, 2013).

The deck can have constant or varying depth. Construction and deck positioning require simple
and inexpensive equipment. Rotation does not allow construction of two parallel bridges, as the
presence of the rst bridge prevents the construction of the second one. The casting yards are
parallel to the obstruction to overpass and contain the entire half-bridge prior to rotation. The pivot
piers cannot be very close to the obstruction, which often results in a longer central span. When the
bridge crosses a channel, additional restrictions derive from building the bridge over the banks.

The negative bending requirements of deck rotation t well with all types of structure compatible
with balanced cantilever construction. Rotation has been applied to constant- and varying-depth
box girders made of steel or PC, and to cable-stayed bridges. The PC box girder of the Ben-Ahin
Bridge in Belgium was cast on temporary supports on a strip of land between two channels, sus-
pended from the tower and rotated to the nal alignment (Figure 2.5). Lightweight concrete was
used for the 168 m main span, and normal concrete for the 128 m anchor span.

21

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


Bridge Launching

Figure 2.5 Ben-Ahin Bridge. (Reproduced with permission from ASCE)

When the main span is in the range 3060 m, the deck is a single-cell, constant-depth PC box
girder. A box girder offers higher structural efciency and easier inspection than a ribbed slab,
and the lower centre of gravity increases the negative bending capacity of the cross-section. On
spans longer than 6070 m, varying-depth box girders are less expensive and aesthetically more
pleasing, although more complex to cast. A 120 m main span was built over the Leie River in
Belgium by rotating deck and pier over the foundation of the pivot pier (de Boer, 2011).

Deck construction is similar to precast segmental construction. Box girders are built using two-phase
long-line match-casting techniques; the bottom form table of the casting cell is supported on scaffold-
ing and is lifted progressively in varying-depth decks. The ribbed slabs are cast in one stage in a
simpler casting cell comprising a central tunnel form and two side forms for webs and wings.

The segments are 47 m long and may be cast symmetrically from the central pier segment towards
the ends, or directionally from one end to the other. Directional casting is used in constant-depth
decks and requires only one casting cell. Most constant-depth decks and all varying-depth decks
are cast symmetrically from the pier segment (Rosignoli, 1998a).

The deck is propped during construction, and prestressing is applied to the completed structure
before rotation. Without the need to anchor the cantilever tendons at each joint and to tension
them as construction proceeds, the tendons can be very long. Prestressing may be internal, external
or mixed. Continuity tendons may also be internal or external.

The centre of mass of the deck is often eccentric from the central pivot point, and a balancing
frame is therefore applied to the rear deck end to control overturning. In most cases, each half

22

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


Bridge launching

of the deck is a simply supported beam with a long cantilever during rotation. The pivot pier resists
most of the vertical load, and the balancing frame resists a low vertical force that is intentionally
created or controlled to provide safety from overturning. The rear support frame slides on a
circular runway beam during rotation.

In a three-span bridge, the optimum length of the end spans is 5560% of the main span for
varying-depth decks and 6570% for constant-depth decks. The deck has end diaphragms, and
longitudinal equilibrium is rarely critical during rotation. If necessary to avoid decompression
of the end supports in service, lightweight concrete may be used in the main span, or the box cell
may be lled with concrete or sand in the end spans, or may have an oversized cross-section.
Temporary counterweights may also be used during rotation. At the end of rotation, the perma-
nent bearings are inserted at the pivot piers after deck lifting and removal of the rotation bearings.

2.3. Jacking and skidding


Jacking and skidding involve deck displacement without major changes in the structural congur-
ation. Incremental launching and rotation are construction processes, and, as such, they involve
changes in the structure on completion of displacement. Jacking and skidding are applied to the
completed bridge and offer more freedom in the design process. The deck must not necessarily be
designed for being moved, and jacking and skidding are therefore also used for demolition of
existing bridges built using different techniques.

Jacking and skidding are trafc-friendly techniques for moving heavy structures in, under and over
active infrastructures. These techniques are not new, but their application has become more and
more frequent for safe and rapid replacement of existing structures with minimal disruption to
trafc and minimal risks to workers and the public. The deck may also include side rails, paving
or ballasted track for immediate opening to service.

Jacking and skidding are applied to many types of structures, including bridges, hangars and roofs
(Jing, 2004). Hydraulic jacking systems with almost unlimited capacity are available for the verti-
cal movement of structures. For horizontal movement, skidding beams with PTFE sliding surfaces
are used in combination with polished stainless steel for the sliding track and hydraulic cylinders to
drive the movement. Jacking and skidding are typically combined for deck lifting, displacement
and lowering onto the permanent bearings. Replacing an existing deck may take 1648 hours
(de Boer, 2011).

Construction of a new bridge by lateral skidding involves casting the deck alongside its nal
position, or incrementally launching the deck into that position and then shifting the deck laterally
into position. This construction method offers rapid construction of short bearing-supported
decks, and several additional advantages (de Boer, 2011).

g The deck static system is the same during construction and in service, and no temporary
stresses arise during displacement. Skidding has minimal impact on the structure, and may
be used to displace structures that were not designed to be displaced.
g When parallel decks are cast in-place, the same falsework can be used for both decks. This
may save substantial time and labour, especially when the falsework is expensive and can be
used twice with minor adjustment.
g The falsework for in-place casting, or the casting yard for incremental launching of the
deck, are located alongside the nal position of the bridge. When the new deck replaces an

23

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


Bridge Launching

existing bridge in service, the interference between the construction site and the existing
bridge is minimised.
g Jacking and skidding involve minimal costs in structures designed for maintenance of
bearings. Jacking is designed using the same criteria as for bearing replacement, and
skidding may be used for both construction and demolition of the deck. For
demolition, the deck is jacked, a steel underbridge is launched longitudinally beneath
the deck from one abutment, and the deck is lowered onto the underbridge and
skidded backwards for progressive demolition behind the abutment. This demolition
method was used for a PC deck over Motorway A86 in France weighing 1800 tonnes
(de Boer, 2011).
g A steel underbridge may also be used to support a full-length casting cell for in-place
casting of the deck at a higher elevation. After completion, the deck is jacked, the
underbridge is launched back to the abutment, and the deck is lowered into position.
This construction method is suitable for short PC or composite bridges over active
highways and railroads, especially when the spans are skewed or have different lengths
or varying depths. A four-span, 250 m overpass weighing 4250 tonnes was built in this
way in The Netherlands (de Boer, 2011).

Bridge 185 of the TGV high-speed railway in France is an example of double utilisation of the
same falsework. Two varying-depth box girders were cast on falsework on the alignment of the
second bridge. On completion of the rst deck, the falsework was lowered and the deck was
skidded laterally to clear the falsework for casting of the second deck. During skidding, the deck
was supported on concrete blocks equipped with a bottom stainless-steel plate sliding over
Neoon (neopreneTeon) plates. The skid force was applied by prestressing bars at the
abutments.

For replacement of the Azergues Bridge 156 on the A6 Highway in France, two new PC decks
were cast on falsework alongside the highway bridge, one on either side. With trafc on only
one of the existing bridges, the deck of the second bridge was demolished and replaced with the
new deck. With trafc on the new deck, the same operations were performed on the second
bridge.

In the presence of two parallel decks, skidding combined with incremental launching makes the
most of casting yard and temporary piers (Figure 2.6). The rst deck is launched along the nal
alignment of the second deck, and is then shifted laterally to the nal position to clear the launch
alignment for the second deck. This technique was used for the two 200 m box girders of the
Tiziano Bridge in Italy, each weighing 3260 tonnes (Figure 2.7). The rst box girder was supported
on skidding shoes during launching and skidding. Skidding at the abutments and the three
permanent piers was driven by prestressing bars anchored to pier brackets at the two outer piers
(Rosignoli and Rosignoli, 2007; Rosignoli, 2008).

2.4. Incremental launching


Incremental launching is a segmental method of deck construction. Full-span launching is applied
to PC spans ranging between 20 and 6070 m, and to steel spans amply exceeding 100 m. Most
incrementally launched PC decks are cast in-place in a xed facility built at grade behind one
of the abutments. The plan and vertical alignment of the casting yard coincide with the local
projection of the launch lines of the deck. The deck is built using repetitive casting procedures, and
is launched as a continuous beam after completion of the cross-section.

24

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


Bridge launching

Figure 2.6 Launch of the right box girder on the temporary piers of the left box girder

Figure 2.7 Skidding of the right box girder into position

25

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


Bridge Launching

Figure 2.8 Integrated-platform underslung MSS. (Photo: Strukturas)

PC decks are mostly launched on 4060 m spans, and incremental launching therefore competes
with span-by-span in-place casting using movable scaffolding systems (MSSs) and with span-by-
span erection of precast segments using self-launching gantries (Rosignoli, 2013). Compared with
span-by-span in-place casting with an MSS (Figure 2.8), incremental launching offers several
advantages:

g Casting the deck in a sheltered facility behind the abutment enhances the safety of workers,
minimises the interference with the area under the bridge, and diminishes the construction
costs, as logistics does not follow production.
g An MSS for 5060 m spans may weigh 10001300 tonnes or more. Shipping, site assembly
and nal decommissioning increase the construction cost of the bridge and lengthen the
time taken for construction. The investment for setting up a casting facility for a launched
bridge may be 1520% of the investment for a full-span MSS, and the labour demand for
assembly and operations is also lower.
g In a launched bridge, prestress is applied uniformly and progressively. In-place casting with
an MSS requires application of prestress as early as possible in order to release the MSS
and to shorten the cycle time. Full prestress application at short curing increases the time-
dependent losses of prestress, and creep of concrete may generate irreversible deck cambers
that disturb drivers.
g A launched deck is cast in a stiff form supported on the ground, while the casting cell of an
MSS is more exible. Form cambering controls the nal span geometry, but load
deections and residual support action at the application of prestress may cause deck
cracking and steps at the construction joints (Rosignoli, 2013).

26

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


Bridge launching

Figure 2.9 Self-launching gantry for span-by-span precast segmental erection. (Photo: VSL
International Ltd)

g Span casting and application of prestress may damage the bridge bearings. The thermal
deformations of a longer and longer continuous beam (the starter abutment is typically the
xity point of the deck) complicate bearing adjustment and may damage the MSS. In a
launched bridge, the bearings are applied on launch completion, and pre-setting is also
easier.

The main advantages of an MSS are the adaptability to irregular bridge geometry and to very
long bridges. The spans are often cast in a weekly cycle, and the erection rate may therefore be
twice the productivity achieved with incremental launching. The presence of only one construction
joint in every span is another advantage.

The number of construction joints is one of the weak points of precast segmental construction.
Compared with span-by-span erection of precast segments with a self-launching gantry
(Figure 2.9), construction by incremental launching offers several advantages:

g In a typical 50 m span, a launched deck has two joints and a precast segmental deck with
3 m segments has 16 joints. Construction joints are weak points in any PC structure. The
construction joints of most launched bridges are located at the span quarters and are
subject to minimal permanent bending. The joints are treated to enhance adhesion and are
never separated; launch prestressing enhances the surface bond further, and through
reinforcement in the joints further enhances control of edge stresses.
g The launch tendons extend over two or three half-span segments, and the external
integrative tendons extend over two or three spans. The tendons of span-by-span precast
segmental bridges are as long as the span in order to release the erection gantry as soon as
possible and thus shorten the cycle time; shorter tendons increase the labour requirements

27

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


Bridge Launching

and the cost of prestressing. Launched bridges are also designed for partial prestressing
during launching, while most design standards prevent decompression of epoxy joints.
g A self-launching gantry for highway segments and 4050 m spans may weigh more than
500600 tonnes. Like an MSS, a gantry requires shipping, site assembly and nal
decommissioning. Setting-up a precasting facility and providing special means of
transportation involve additional investments and nancial exposure, which require long
bridges for cost depreciation.

Rapid deck construction is the major advantage of segmental precasting. A self-launching gantry
may erect a 50 m span with epoxy joints in 23 days, while incremental launching of half-span
segments takes 2 weeks. For short bridges this advantage is only apparent, as segment erection must
be delayed until most of the piers have been completed, so as not to interrupt the fast operations of
the gantry. Gantry erection is also compatible with irregular deck geometry.

2.4.1 Structural configuration


With the progress of materials technology and calculation methods, the continuous beam struc-
tural system has shown several advantages over statically determinate schemes. Continuous beams
require less structural material, and ensure smaller deections and better control of fatigue than do
the simple spans. Avoiding solutions of continuity improves the seismic response of the bridge,
and the smaller number of bearings and expansion joints decreases the maintenance costs. Even
in the event of differential settlement of supports, the hyperstatic stresses are reduced by creep
of concrete, and may be corrected by shimming.

When the bridge length does not cause excessive dimensional variations due to temperature and time-
dependent effects, the structural advantages of a continuous beam suggest permanently maintaining
the multispan continuity used during launching. In this case, on launch completion it is necessary
only to replace the launch bearings with the permanent bearings and to complete prestressing.

Long bridges and high thermal differentials may suggest dividing the deck into shorter continuous
beams. In the 49.16 Lot of the high-speed railway TGV Atlantique in France, a 445 m PC box
girder was launched as a continuous beam. On launch completion, the temporary prestressing
of four expansion joints was released to divide the deck into ve three-span continuous beams,
having spans of 29.2, 30.6 and 29.2 m, which were jacked into position on the launch bearings.

The same technique was used in South Africa on the Olifants River Bridge, which includes 23 PC
railway spans, each 45 m long, to give an overall length of 1035 m. The nal structural system
consists of two 11-span continuous beams xed at the abutments and a simply supported
expansion-joint central span. This scheme applies the traction/braking loads to the abutments and
permits the use of slender piers. The 23 spans were launched as a single continuous beam. On
launch completion, the deck was permanently xed to the launch abutment. After relieving the
launch prestressing of the rst expansion joint, the 12-span continuous beam was jacked forward
to move the central span into position. After opening of the second joint, this operation was
repeated on the 11-span front section, which was xed to the opposite abutment.

A similar solution was adopted in France for the highway Oli Bridge, composed of 15 PC spans,
each 41 m long, with a longitudinal 5.4% gradient. The piers were 60 m tall, which dictated
downhill launching and a nal design of two continuous beams xed at the abutments, with a
central expansion joint. On launch completion, the upper box girder was permanently xed to the

28

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


Bridge launching

launch abutment, and progressive release of the temporary tendons in the central joint was used
for braking of the lower girder during sliding into position.

Simple spans are often adopted in railway bridges because of the low cost of sub-ballast expansion
joints and to minimise railstructure interaction. The Sinntal Bridge for the ICE high-speed rail-
way in Germany includes eight 42.5 m PC spans, a 51.5 m span and a 30.8 m span, and was built
by launching as a continuous beam a sequence of simply supported box girders temporarily con-
nected to each other by joint prestressing. On launch completion, the spans were disconnected and
progressively jacked into position. The central movable span of the Basrah Bridge in Iraq (Seifried
and Wittfoht, 1979) was also launched together with the access spans as a single continuous beam,
and released on launch completion.

The structural conguration used for launching may also be modied by increasing the grade of
redundancy. If a curvature transition is located at the centre of a low-level bridge, an aerial casting
cell may be created at the centre of the bridge to launch two deck sections one after the other in the
opposite directions. Construction duration is unaffected and the thrust force halves, which may be
imperative for PC bridges longer than 8001000 m. This scheme has also been applied to steel
bridges, where light portal cranes are used to serve an elevated assembly platform.

Two continuous beams may also be launched from the opposite abutments and connected on
launch completion at the centre of the bridge (Frizzi and Giovannini, 1977). When the deck
includes two distinct sections and each section has launchable geometry (a rectilinear section and
a circular section, or two sections with opposite plan curvature for a skewed S-crossing of a river,
Figure 2.10), launching from both abutments is compatible with a non-launchable general bridge
geometry (Geier, 2010). Launching from both abutments doubles the cost of the casting facilities
but shortens construction duration and diminishes the thrust force needed for launching. In the
nal phases of launching, the launch nose is raised to slide over the front diaphragm of the deck
section already in place. A temporary pier under the joint segment facilitates casting of the closure
pour, but good results may be also achieved with the midspan closure techniques used for
balanced cantilever bridges.

The Schnaittach Bridge in Germany was built by launching a 424 m PC deck section with constant
plan curvature downhill from the upper abutment, and an 864 m deck section including a front
clothoid spiral and a rear rectilinear section from the lower abutment. Casting the front clothoid
section required progressive rotation of the casting cell to create angle breaks between the deck
segments. Lateral shifting of the launch bearings was necessary to accommodate the irregular web
alignment in the clothoid section (Figure 2.11). On launch completion, the two deck sections were
connected with in-place closure.

A similar solution was used for the Schrotetal Bridge in Germany. The 495 m composite box
girder was built by launching a 404 m steel U-girder from one abutment and a 51 m thinner U-
girder from the opposite abutment. A 40 m varying-depth segment was assembled on the ground
and strand jacked into position to close the steel girder, and the concrete slab was cast segmentally
with a forming carriage.

Many steel girders have been launched from the opposite abutments. This method is used for long-
span portal frames with inclined legs and a steel orthotropic deck (Matildi and Matildi, 1990), and
for continuous beams with a concrete slab (Bernard, 1997). Launching continuous beams from the

29

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


Bridge Launching

Figure 2.10 Auenbach Bridge. (Photo: Schimetta)

Figure 2.11 Launch bearings on skidding shoes. (Reproduced with permission from ASCE)

30

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


Bridge launching

opposite abutments also permits overtaking a long central span. Temporary masts and stay cables
were used to support the cantilevers of the central span of the Mainingen Bridge in Germany
prior to midspan closure. When the bridge includes one long central span and approaches
compatible with full-span launching, incremental launching construction from the opposite abut-
ments with midspan closure is often faster and less expensive than balanced cantilever construction.

The exibility of long steel cantilevers causes large angle breaks in midspan, which must be
recovered prior to closure. Several methods are available to align the cantilevers. The box girders
of portal frames are isostatic during launching, which simplies rotation about the top of the inclined
legs. The front region of a continuous beam may be made temporarily isostatic by releasing a eld
splice in the rear span. The cantilevers of the 147 m main span of the Bayindir Bridge in Turkey were
realigned by releasing eld splices in the rear 73.5 m spans. The two girders were rotated about the
main piers up to midspan alignment and closure, and the eld splices in the rear spans were then
jacked back into position (Popov and Seliverstov, 1998). After alignment, the tips of the cantilevers
are cut and prepared for welding from a suspended scaffold (Zhuravov et al., 1996).

The increase in redundancy may be more substantial. A steel or PC deck can be launched over an
arch and permanently framed to the spandrel columns, or can be launched over temporary piers
and suspended from one or more towers to obtain a nal cable-stayed bridge; both solutions have
been tested with steel and PC decks. A deck launched over temporary piers can also be used to
establish a working platform for assembly of the ribs of a tied arch with strand-jacking towers
extending from the launch piers. After completion of the arch ribs, the deck is suspended from the
arch, and the jacking towers and the launch piers are removed.

2.4.2 Final articulation and seismic design


On launch completion, the deck is a continuous beam supported on low-friction bearings. This
conguration is modied to achieve the permanent static system of the bridge, and several alterna-
tives are possible.

g Poly-tetrauoroethylene (PTFE, Teon) bearings are used at the piers and the end
abutment, and the launch abutment provides xity and resists most of the longitudinal
loads. This scheme may be improved with dampers at both abutments and sacricial shear
keys at the launch abutment. Failure of the shear keys releases the deck during the seismic
design event, the seismic demand is shared between the abutments, the piers resist
additional load via sliding friction or pier dampers, and the equivalent system damping
increases. This scheme is devoid of re-centring capability.
g The deck is xed at both abutments, an expansion joint is created at the centre of the
bridge, PTFE bearings are used at the piers, and the two abutments resist the longitudinal
loads applied to the tributary deck sections.
g Fixed bearings are used at the tallest piers, and PTFE bearings and dampers are used at the
other piers and the abutments (Llombart and Revoltos, 2000). This scheme improves the
distribution of the seismic demand between piers and abutments and increases the
equivalent system damping. Steel or reinforced-concrete (RC) pier-cap shear keys may be
used at the xity piers to achieve pier-base plastic hinging mechanisms (Tegou and Tegos,
2012). This scheme may be too demanding in railway bridges as the braking/traction forces
are applied to the restraint piers.
g The deck is made structurally continuous with some piers to create plastic hinging
mechanisms at the top and bottom of the piers (Calvi et al., 1996).

31

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


Bridge Launching

g The deck is isolated from piers and abutments by low-damping rubber bearings, lead-plug
bearings, friction pendulum systems and isolation/dissipation devices (AASHTO, 2000;
Kelly and Naeim, 1999).

In the transverse direction, the deck is isolated or rigidly connected to piers and abutments. When
the lateral guides of PTFE bearings are insufcient for transfer of lateral seismic response forces,
sliding shear keys are installed between the deck and the pier caps to allow longitudinal displace-
ments (Abeysinghe et al., 2002).

2.5. Launching over arches


The deck of a concrete arch may be an open grid of simply supported precast I-beams, a continu-
ous steel girder, or a continuous PC box girder or ribbed slab. Precast I-beams and steel girders
require massive air work during construction. Beam launchers are typically used to erect precast
I-beams, while cable cranes may have enough capacity for steel girders. Cable cranes or motorised
trolleys rolling along the steel girders are used for delivery and placement of precast deck panels,
and forming carriages or conventional shoring systems may be used for in-place casting
(Rosignoli, 2013).

A continuous PC box girder offers higher exural and torsional stiffness, a smaller number of
bearings, no internal expansion joints, and better structural response by providing additional
stiffness at the deck level. The structural behaviour is similar to that of a deck-stiffened arch, with
reduced bending at the arch springings that allows for more slender arch ribs. A continuous
PC ribbed slab offers similar advantages and is easier to cast, but the deck is more exible and
maintenance is more complex.

A continuous PC box girder can be built by incremental launching, by span-by-span in-place cast-
ing using an MSS, and by span-by-span erection of precast segments using a self-launching gantry.
One of the main advantages of incremental launching is the possibility of increasing the spacing
between spandrel columns and approach piers without the increased cost and weight of an
MSS or a self-launching gantry as the span increases. Optimisation of the deckarch interaction
and the choice between curved or polygonal arch ribs are thus less restrained.

Continuous decks incrementally launched over arches are a brilliant solution to the need to ensure
the safety of workers and to achieve a short duration of construction and cost savings. The Neckar
River Bridge in Germany was built by launching two 15 m wide, 365 m long PC box girders over
two parallel arches and approach spans (Figure 2.12). The arches have a span of 154.4 m and a
rise of 49.9 m. A similar solution was adopted in the Isere River Bridge in France (Placidi and
Virlogeux, 1991). An 8.6 m wide, 234 m long PC ribbed slab was incrementally launched over a
123 m arch and its approach spans; the arch has a rise of 24 m.

The deck is launched from one abutment towards the other, and the arch must have enough
exural stiffness to resist the asymmetrical loads. Span-by-span casting with an MSS and erection
of precast segments with a self-launching gantry are also directional construction processes,
and bending in the arch ribs is higher because of the additional weight of the construction
equipment (Figure 2.13). Launching the deck symmetrically from the opposite abutments could
avoid load asymmetry; however, symmetrical loads on only some of the spandrel columns would
also induce bending in the arch, and the cost of two casting facilities rarely makes this solution
competitive.

32

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


Bridge launching

Figure 2.12 Launch over arches for the Neckar River Bridge. (Reproduced with permission from ASCE)

Bending in the arch ribs due to the frictional loads of deck launching is minimised by the use of
temporary stay cables that connect the spandrel columns to the launch abutment. The stay cables
also diminish the net launch reaction applied to the abutment foundations. The stay cables are
usually anchored to the spandrel columns, so that the load on the opposite sides of the columns
may be different. The stay cables do not increase the construction cost much as the same strand
and anchorages are used during cantilever construction of the arch and deck launching.

Figure 2.13 Underslung MSS for the Wumbach Viaduct. (Photo: DB AG)

33

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


Bridge Launching

Figure 2.14 Veitshochheim Bridge for ICE high-speed railway. (Reproduced with permission from ASCE)

In the Veitshochheim Bridge for the ICE high-speed railway in Germany, temporary stay cables
were integrated with counterweights to control launch bending in the polygonal arch (Figure 2.14).
The dual-track box girder, 1260 m long and weighing 42 500 tonnes, is one of the heaviest PC
decks ever launched (Flugel, 1987; Leonhardt, 1991; Theiner, 1987; Zilch, 1987). The arch spans
162 m with a rise of 24 m, and the approach spans have a constant length of 53.5 m.

Filling the arch ribs with water and progressively emptying rib sectors at the arrival of the PC deck
on the spandrel columns has been studied for the 300 m arch of the Hoover Dam Bridge in the
USA, but never attempted. During deck launching on the approach spans, the arch ribs are lled
with water. When the arch is full, the load generated by the water is similar to the load applied by
the deck at the end of launching (Figure 2.15). The deck is launched to the rst spandrel column,

Figure 2.15 Arch stabilisation by means of water filling

34

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


Bridge launching

and water is pumped out from the springing sector of the arch, so that the weight of the deck
replaces the weight of the water. This process is repeated until the deck reaches the approach spans
on the opposite side of the arch.

2.6. Launch of cable-suspended decks


Numerous cable-stayed bridges have been designed with two planes of stay cables supporting a
steel grillage comprising edge girders, oorbeams and substringers and completed with a PC deck
slab. Precast deck panels are made continuous with the steel grillage with in-place stitches over the
top anges of oorbeams, substringers and edge girders. Built-up I-girders are used for the edge
girders of single-deck bridges, and deep trusses are used for double-deck bridges because of clear-
ance and ventilation requirements. The stay cables are anchored over the edge girders or in anchor
pipes bolted to their outer face.

A composite grillage is much lighter than a PC deck, and this generates cost savings in stay cables,
towers and foundations. The deck is erected with oating cranes or deck-supported derrick
platforms that handle modules of edge girders and oorbeams and precast deck panels individu-
ally. Light erection equipment and small load imbalance diminish the transient stresses of staged
construction in towers and foundations. The drawbacks are the cost of the steel grillage, its
maintenance cost over time, the need for an efcient temporary restraint between the deck and the
tower during construction, and the need for a wide deck to ensure lateral stability of long
cantilevers during cantilever construction.

PC ribbed slabs comprising edge girders and oorbeams are also used in the cable-stayed decks.
The stay cables resist most of the negative bending that characterises balanced cantilever con-
struction, and the use of multiple stay cables with closely spaced anchor points diminishes the
demand for exural stiffness in the deck with both harp and fan arrangements. A ribbed slab is
easier to cast than a box girder, the stay cables are anchored at the bottom of recess pipes
embedded in the edge girders, the latter directly resist the longitudinal component of the pull in
the stay cables, and maintenance is less expensive than for a composite grillage. However, a PC
ribbed slab is heavier than a composite grillage, load imbalance is larger during staged construc-
tion, and two planes of stay cables are still necessary because of the poor torsional constant of the
open section.

Several cable-stayed decks have also been designed with PC box girders. Box girders typically have
a single-cell section, their width may reach 1820 m, and ribs, diagonal struts and combinations of
ribs and struts have been successfully used to widen the top slab further. The torsional strength
and stiffness of a wide box girder integral with a central pylon are typically sufcient to support
the deck with a single central plane of stay cables. Many cable-stayed bridges carrying separated
highways have been designed with a single plane of cables to simplify the tower, to diminish the
number of anchorages, to improve the aerodynamic stability of the deck with a streamlined prole,
and to diminish the drag coefcient.

Cable-stayed box girders are made of steel or PC, or are of composite construction. Streamlined
steel box girders erected with oating cranes or lifting frames are used for the longest spans
because of the higher strength-to-density ratio of steel. PC is the typical choice for shorter spans
due to the lower cost of materials and the less maintenance needed over time. Single-cell PC box
girders supported by a central plane of stay cables are too wide and heavy for segmental precast-
ing, and are mostly cast in-place using form travellers (Rosignoli, 2013). Prestressed composite box

35

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


Bridge Launching

girders with steel corrugated-plate webs are earning popularity in the 100200 m span range
because of weight saving, high exural efciency, and the web capability of not interfering with
post-tensioning and the longitudinal component of the pull in the stay cables.

Compared to a PC ribbed slab, the constant of torsion of a box girder is 23 orders of magnitude
higher, the moment of inertia is one order of magnitude higher and the cross-sectional area is
similar. A box girder is, therefore, perfectly suitable for incremental launching over temporary
piers and suspension from the towers on launch completion. When the area under the bridge
allows partial disruption during construction, deck launching over temporary piers offers several
advantages:

g Construction is faster and less expensive than with balanced cantilever erection. Approaches
(Jiang and Yang, 1998) and main span can be erected with one learning curve, less
investment in special equipment, and simpler logistics for deck construction and fabrication
of the stay cables. When a steel main span is combined with PC approaches, a portion of
the steel span can be used as a launch nose during incremental launching of the approach
spans (Figure 2.16) (Lockmann and Marzahn, 2009).
g The towers can be erected out of the critical path during deck launching.
g On launch completion, the deck is used as a working platform for cable fabrication.
Materials are delivered along the deck by conventional means of ground transportation,
and heavy cranes may be operated on the deck. The number of tensioning operations
diminishes, and control of geometry and the pull in the stay cables is much simpler.
g The deck is unaffected by aeroelastic disturbance during construction.

Figure 2.16 Steel deck segment used as a launch nose for the right approach spans. (Photo:
Strassenbau NRW)

36

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


Bridge launching

Figure 2.17 Launch of the first deck segment of the Palizzi Bridge

When the deck is over water, the drawbacks include the number of temporary piers and the inter-
ference with the navigation channel. The 21.8 m wide, 527 m long PC box girder of the Wandre
Bridge in Belgium was launched over temporary piers placed in the Meuse River and a parallel
channel. With 18 m long deck segments, launching the 11 800 tonnes of the deck required a
35 m launch nose. On launch completion, the deck was suspended from a 95.5 m tall A-tower
to attain the nal static system with two cable-stayed spans of 144 m and 168 m (Greisch, 1993).

Launching a low-level deck over temporary piers is particularly advantageous when the area under
the bridge can be partially disrupted. In wide railway crossings, some tracks may be temporarily
deactivated, or their spacing may be compatible with foundations for the temporary piers. The
dual-track LRT Palizzi Bridge in Italy was launched over six electried railway tracks with the help
of two temporary piers (Rosignoli, 1998d; Martinez Y Cabrera and Rosignoli, 2001). The PC deck
is 156 m long and 1.0 m deep, and the main span is 66 m long (Figure 2.17).

When no temporary piers can be used for launching, the pylon may be made integral with the deck
and used to deviate temporary cables that support the front cantilever during full-span launching.
This construction method was used for a pedestrian bridge located on the top of an 80 m tall
intake tower within a reservoir close to Granada in Spain (Llombart and Revoltos, 1996).

Multispan cable-stayed bridges with integral pylons are optimal candidates for low-level launch-
ing over railways. The 23.8 m wide, 580 m long composite deck of the Coast Meridian Overpass in
Canada was launched over 50 tracks (Figure 2.18). The deck includes four 30 m tall steel pylons, a
single central plane of stay cables, and ve cable-stayed spans of lengths ranging from 111 to

37

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


Bridge Launching

Figure 2.18 Coast Meridian Overpass. (Photo: Somerset/KWH)

125 m. The deck framing system comprises one 3 m deep U-girder on either side of a central pair of
spine beams, which are aligned with pylons and stay cables and connected to the U-girders with
full-depth diaphragms. During full-span incremental launching, the steel frame was supported
under the U-girders, and lead pylon and stay cables were used to support the 125 m front canti-
lever in combination with a long launch nose (Gale, 2011).

Multispan cable-stayed bridges with integral pylons are also optimal candidates for high-level
launching, as the deck establishes a working platform for the activities to be performed above the
deck on launch completion. The 32 m wide, 2460 m long streamlined steel box girder of the Millau
Viaduct in France includes seven 87 m tall inverted-V steel pylons and a single central plane of stay
cables. The length of the eight cable-stayed spans is 204 m for the end spans and 342 m for the
interior spans. The deck was launched from the opposite abutments with the help of one tempor-
ary pier per span. The lead pylons were used to anchor some of the permanent stay cables to
support the front cantilevers during launching. The rear pylons were delivered on the deck, rotated
to vertical, and completed with the stay cables on launch completion (Virlogeux et al., 2005).

The streamlined steel box girder of the 260 m main span of the self-anchored suspension
Hangzhou Jiangdong Bridge in China was also launched over temporary piers (Zhang et al.,
2010). Because of the cambered prole of the deck, the launch bearings were adjusted vertically
during launching. The stiffening girder of a self-anchored suspension bridge must be completed
prior to application of the hangers, and temporary piers are therefore necessary anyway.
Incremental launching of the 47 m wide deck simplied construction and avoided the use of oat-
ing cranes.

A permanent deck launched over temporary piers can also serve as a working platform for rib
erection of a tied arch (Figure 2.19). This solution was adopted for the 218 m span of the
Reggio Emilia Bridge in Italy (Rando et al., 2010). After launching the 27 m wide single-cell steel
box girder complete with the orthotropic deck, the three temporary piers were extended over the
deck to serve as strand-jacking towers for the arch rib segments. After closure of the arch rib and
removal of the strand-jacking towers, locked-coil strand hangers were applied to suspend the box

38

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


Bridge launching

Figure 2.19 Removal of the temporary piers at the end of construction

girder from the arch and remove the launch piers. A similar solution was used for the arch bridge
over the River Loire in Orleans, France (Hoeckman, 2001).

2.7. Geometry constraints of incremental launching


Incremental launching is a very versatile construction method, and has a wide range of appli-
cations. Limited at rst to the construction of bridges with simple geometry, it has since been used
successfully for the construction of structures having increasing geometric complexity (Favre
et al., 1999).

During launching, the deck is a continuous beam supported on launch bearings and restrained by
lateral guides. Misplacement of bearings and guides causes hyperstatic stresses in the deck and the
piers and accelerated wear of the launch systems. Bearings and guides are therefore tightly aligned
with the surfaces of the deck and the launch nose they will come in contact with during launching.
This requires that a common geometry is set up for the casting cell, launch nose, launch bearings
and lateral guides.

The allowable geometries can be dened mathematically by considering the production of identi-
cal segments (Ontario, 2006). For a rigid body (deck and launch nose) to slide within another rigid
body (the launch alignment provided by launch bearings and lateral guides), the solid must be

39

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


Bridge Launching

superimposable onto itself by translation, rotation or rotationtranslation. The only lines that are
superimposable by rigid displacement are the rectilinear segment (translation), the arc of the circle
(rotation) and the circular helix, which combines rotation and translation through the pitch of the
helix. A sequence of individually launchable lines is not launchable if the rigid displacement of the
sequence generates non-overlapping areas (AFGC, 1999). Launching from the opposite abut-
ments simplies the geometry requirements, as different launchable lines may be used for the two
sections of the deck.

Transverse deck stability requires two support lines during launching, which are parallel in recti-
linear bridges and concentric in circular bridges. The launch bearings are located under the webs,
and the two support lines must be launchable as a whole to avoid torsion and distortion of the
cross-section. Even if the longitudinal axis of the deck is launchable, therefore, cross-fall tran-
sitions at the bottom slab level may make the deck non-launchable.

The bridge designer starts from a plan layout of the bridge, a longitudinal prole and a law of
cross-fall variation in the area of the bridge. The longitudinal prole is a representation of the
elevation developed on the vertical cylinder containing the deck axis. The plan layout is a deck
projection on the horizontal plane, which is not the real three-dimensional curve of the deck if the
latter is not horizontal. Finally, cross-fall transitions relate to deck extrados, the launch lines relate
to the soft, and the cross-section may be distorted to make cross-fall transitions compatible with
the launch requirements.

When the deck carries two-way road trafc, the two halves of the top slab often have opposite
cross-fall, and the deck soft is a horizontal line orthogonal to the longitudinal plane of symmetry
of the deck. This is the typical conguration also for railway bridges carrying one or two tracks,
with both ballasted track and direct xation. When the deck carries one-way highway trafc, three
solutions are possible to handle cross-fall:

g The cross-section is kept symmetrical and rotated to cross-fall. Extrados and intrados of
PC solid and voided slabs are kept parallel for aesthetic reasons, and the deck soft is
therefore inclined in the cross-section plane (De Clercq and De Ridder, 2003).
Launching generates lateral drift forces, and the launch guides are designed accordingly.
Box girders are used on longer spans, the lateral drift forces would be excessive, and the
soft is modied so as to create two horizontal launch surfaces at different elevations
under the webs.
g The cross-section is made asymmetrical by keeping the bottom slab horizontal and by using
webs of different depth. This solution is not recommended in a PC box girder because of
the absence of symmetry for reinforcement and post-tensioning, the different edge stresses,
and the risks of errors during design and construction (AFGC, 1999). This, however, is the
standard solution for the launch of the steel U-girder of a composite box girder.
g The bottom slab and webs are kept symmetrical, the bottom slab is kept horizontal, and
the top slab thickness is adjusted to cross-fall. This solution is also not recommended
because of the additional weight and the structural asymmetry of the cross-section. This
solution has some merits for local cross-fall transitions at the ends of the bridge.

In relation to the general deck geometry, the incremental launching method can be used for
rectilinear bridges or where the deck has a curve of constant radius throughout its length. The
longitudinal axis of the deck may be, in order of increasing complexity:

40

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


Bridge launching

1 Rectilinear in plan and with constant longitudinal gradient. The launch lines are rectilinear,
parallel, and on different elevations if the deck soft has cross-fall. The deck is launched by
pure translation.
2 Rectilinear in plan and circular in prole. The deck is launched by rotation along a circular
cylinder with a horizontal axis, and the two launch lines are parallel arcs of a circle. Decks
with a parabolic prole are not launchable; however, the radius of vertical curvature is
often so large that steel plates or hardwood shims may be applied to the web soft to make
the deck launchable. Progressive shimming or jacking of the launch bearings is also
possible, especially for the launch of light steel girders.
3 Circular in plan and horizontal in prole. The deck is launched by rotation along a circular
cylinder with a vertical axis. Launching a deck with varying plan curvature (e.g. a non-
launchable clothoid spiral) requires progressive skidding of launch bearings and lateral
guides, and pier design for the resulting load eccentricity.
4 Circular in plan and with constant longitudinal gradient. The deck is launched by rotation
translation along a helix contained in a circular cylinder with a vertical axis.
5 Circular in an inclined plane (Figure 2.20). The deck is launched by rotation along a
circular cylinder with an inclined axis, and the plan and vertical projections of the deck axis
are arcs of an ellipse. When the soft has cross-fall, the two launch lines are contained in a
attened cone with an inclined axis (Bennett and Taylor, 2002). Large plan and vertical
radii often allow the launch of decks that are circular in plan and prole by inclining the
launch cone. Marked curvatures require considering distortion of the geometry from a
circle to an ellipse when positioning launch bearings and lateral guides (AFGC, 1999;
Giovannini, 1972).

Cases 1 to 4 are geometric degenerations of the general launch-cone case (5). The launch nose par-
ticipates in the deck displacement and should be a geometric extension of the deck. The launch
nose, however, is typically rectilinear to facilitate its reuse in future projects. Although the exi-
bility of the steel nose simplies alignment corrections, a rectilinear nose is a priori not launchable
in a curved bridge, and the geometry irregularities modify the launch stress distribution in the
front deck region.

When the deck has a circular or pseudo-circular vertical prole, the nose may be set: tangential to
the circle; aligned with the chord (nose tip and nosedeck joint both on the circle); or somewhere in
between these two positions. A tangential nose may require vertical adjustment of the launch bear-
ings (shimming with convex prole, and lowering with concave prole) at nose landing and up to
the arrival of the PC deck, to calibrate the reaction provided by the front support. A nose aligned
with the chord minimises the amount of correction required.

Plan curvature may have more signicant consequences, as a rectilinear nose shifts laterally over
the front supports during launching (AFGC, 1999). The launch nose may be made polygonal by
shimming the eld splices between segments, but the segments are typically designed to be as long
as possible to minimise eld splicing, and the geometry correction is therefore very approximate.
The nose may also be aligned with the chord, so that nose tip and nosedeck joint are both on the
theoretical launch alignment, and the launch bearings are widened to allow lateral shifting. Lateral
shifting may be halved by setting the nose parallel to the chord but offset outward by one-half of
the sag of the circle at the centre of the chord. In either case, the nose applies a torsional moment to
the deck, and two planes of lateral bracing are necessary in the nose to provide torsional strength
and stiffness.

41

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


Bridge Launching

Figure 2.20 Launch of the Val Restel Bridge with 150 m plan radius. (Reproduced with permission from
ASCE)

Geometry constraints and uniform launch stresses throughout the length of the bridge suggest the
launch of constant-depth decks. In bridges with progressively longer spans towards the centre,
varying-depth decks have been attained by using a vertical radius of curvature for the deck soft
that is different from the one for the top slab. The Ile Falcon Bridge in Switzerland was launched
with a convex vertical radius of 24 900 m for the top slab and 60 000 m for the soft. Over the
720 m length of the bridge, the depth of the PC box girder was thus increased from 2.1 m at the
abutments to 3.7 m in the middle of the central 73 m span (Favre et al., 1999). The front section of
the Thurrock Viaduct in the UK also has varying depth (Kirk et al., 2005).

Some types of launch bearing can be shifted laterally and adapted to varying plan curvatures and
varying-width decks. However, when designing a launched bridge, the varying-radius transition

42

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


Bridge launching

curves should possibly be located outside the bridge. If a transition curve affects the end section of
the bridge, the soft of a PC box girder can be designed with constant plan curvature, and the side
wings of the top slab can be adjusted to the design alignment, as transition curves begin with small
radius variations.

2.8. Launch techniques


Handling a bridge deck involves large forces and requires the guide and control of big volumes.
The force necessary to launch the deck is proportional to its weight, as both the friction resistance
and the longitudinal force produced by the launch gradient are a function of this force. In the long-
est PC bridges the launch force can amply exceed 10 MN, although in most cases it is only a few
mega-newtons. This is nearly the same strength as prestressing tendons, and in the rst PC
launched bridges it was natural to use the prestressing materials and equipment already available
in the yard. This led to the development of towing devices comprising one or two prestressing jacks
anchored to a foundation and acting on strands or bars anchored to the deck.

Over time, some inconveniences of the tow systems stimulated the development of thrust devices,
some applicable to relatively modest loads, some suitable for higher loads. As a result, the launch
equipment currently available presents a wide range of mechanical characteristics, power and cost
depending on the specialist eld of its utilisation (Rosignoli, 2000a):

g Electro-hydraulic winches pulling reeved ropes are used only for light steel girders, and
their use is progressively being abandoned due to the poor control of movement. When the
main winch is placed between the abutment and the rear end of the steel girder, a counter-
winch is used for backward pulling in case of need. The two winches may be combined into
a capstan to accelerate the operations (Rosignoli, 2013). Launch capstans and roll bearings
may lead to launch velocities of 0.51.0 m/min. Recovery of the nose deection at landing
and repositioning of pulling and braking ropes slow down launching, and it may take half a
day to launch a 5070 m bridge section.
g The least expensive launch systems apply a tow force to strands or prestressing bars
anchored to the deck and to the abutment by means of long-stroke, double-acting, hollow-
plunger cylinders. These launchers are suitable for uphill launching of light loads such as
short PC decks, medium-length steel girders and concrete slabs launched over the steel
girders.
g Intermediate hydraulic launchers apply a thrust force to the rear end of the deck by self-
clamping to reaction beams. These launchers are t for uphill launching of medium loads
such as prestressed composite box girders, long steel girders and medium-length PC decks.
g The most expensive electro-hydraulic launchers apply the thrust force by friction and their
use is generally reserved for the movement of large masses over short times, for downhill
launching and for solving particular control requirements of the launch forces. As friction
launchers are extremely adaptable, when available they are often used also for short
bridges.

Regardless of the transfer modality of the launch force to the deck, every type of launcher requires
an anchor element restrained to the ground that resists the launch reaction. The foundation of the
abutment is the most logical candidate for load transfer. The longitudinal load applied to the abut-
ment during launching is often higher than the permanent design load. The vertical load is also
higher, as the tributary deck length on the abutment is longer during launching than in the nal
structural conguration of the deck.

43

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


Bridge Launching

The most demanding load condition for the launch abutment is often reached at the end of launch-
ing, when the launch reaction is maximum and the vertical load diminishes toward the end support
reaction of the continuous beam. When necessary, the foundation of the casting yard is connected
to the abutment to create a long concrete bed that cooperates by friction, or the abutment is
secured with ground anchors (VSL International, 1977). The abutment may also settle at the
beginning of launching, and shimming is relatively easy as the deck is almost isostatic in these
stages.

If the launch abutment is tall and cannot resist the launch reaction, the thrust devices may be
anchored to a special foundation block located between the abutment and the casting cell. The
foundation block is positioned as close as possible to the abutment to diminish the length of the
casting yard and the cost of the temporary deck extension required for the last launch phases,
when the rear end of the deck is between the launchers and the abutment.

2.8.1 Launch of light superstructures


Light superstructures (short PC decks, medium-length prestressed composite box girders, and
long steel girders) are typically launched with inexpensive equipment, as the launch force is small
and the assembly yard must be set up and dismantled rapidly and at low cost.

Deck towing with prestressing bars or strands is the most common solution, although it presents
some disadvantages. The typical gradient of highway bridges, 14%, is similar to the friction
coefcient of steelPTFE contact surfaces. Therefore, uphill launching with new or well-greased
Neoon pads requires anchoring the deck to the abutment during construction of the new seg-
ments to prevent uncontrolled backward sliding. Downhill launching is not recommended with
inexperienced crews and requires bidirectional launch devices.

The rear thrust devices are based on long-stroke, double-acting hydraulic cylinders that self-clamp
to reaction beams or racks and push the rear end of the deck forward. The extraction rails of the
casting cell for a PC or prestressed composite box girder may be used as reaction beams for the rear
thrust cylinders. For the launch of lighter steel girders, the thrust cylinders are anchored to full-
length steel reaction beams. The rear thrust systems have the same weak points as tow systems
when launching along inclined planes.

2.8.1.1 Tow systems


Coupled prestressing bars and double-acting, hollow-plunger, hydraulic cylinders anchored to the
deck or to the front wall of the abutment have been used many times to launch PC and steel girders
when the direction of the force to be applied is certain. For intermediate launch gradients, tow
systems based on coupled bars are combined with friction locks or antagonist bar systems.

The thrust force is often in the range 0.81.5 MN. Higher forces may be achieved by combining
pairs of bars. The 7.5 MN peak thrust force for the Petra Tou Romiou Viaduct in Cyprus required
the use of six 50 mm bars each having a braking load of 1.96 MN. The bars were anchored to two
vertical thrust beams crossing the slabs of the box girder (Llombart and Revoltos, 2000). Two
pairs of 60 mm bars were used for the launch of PC bridges in Spain.

The short stroke of prestressing jacks suggests the use of anchor boxes lodging two jacks so that
the jacks alternate (Figure 2.21). During the launch stroke of one jack, the other jack returns to
idle to be ready to pull the bar when the rst jack reaches its end of stroke, thus avoiding bar

44

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


Bridge launching

Figure 2.21 Rear 1.3 MN thrust jacks for 0.2 m launch strokes. (Reproduced with permission from ASCE)

relieving. The anchor boxes are windowed for the insertion of guillotine anchor plates acting on
anchor nuts that are advanced progressively along the bar. The hollow launch cylinders are
designed to allow anchor nuts and bar couplers to pass through.

When only one launch cylinder is used on each bar, the pull in the bar is relieved at the end of each
launch stroke. The stroke necessary for re-stressing the bar is lost, and this increases the number of
launch cycles and the launch duration. If the casting/assembly yard is long, bar elongation at ten-
sioning may also result in excessively short effective strokes. Special hollow-plunger cylinders
designed for launching provide launch strokes longer than 1 m (Figure 2.22). Single launch
cylinders are used only for horizontal launches.

The drawbars are placed beneath the deck or alongside, on both sides. When the uphill launch
gradient exceeds 0.20.3%, the deck is anchored during the return stroke of the launch cylinders
to prevent uncontrolled backwards sliding. The drawbars may be used for anchoring. When two
jacks are used on each bar, the bar is always in tension and one jack is always active in the anchor
box. When single launch cylinders are used at the abutment, the drawbars may be divided into
coupled segments that are as long as the launch stroke. When the plunger reaches the end of stroke,
the bar is anchored to the abutment with a guillotine anchor plate, the plunger is retracted, the
superuous bar segment is eliminated, and the new bar is connected to the plunger for a new
stroke. This system is effective only with long-stroke cylinders, and requires a great number of
short bars and bar couplers.

High thrust forces suggest the use of strand cables. In principle, by using strands the thrust force is
limited only by the capacity of the jacks, the addition of further strands being always possible.

45

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


Bridge Launching

Figure 2.22 Long-stroke hollow-plunger cylinder anchored to the abutment. (Reproduced with
permission from ASCE)

However, it is generally convenient not to exceed 2.5 MN in each draw cable to use 19-strand
prestressing jacks for launching. The elasticity of the draw cables is one of the weak points of these
launch systems. When the friction coefcient drops from breakaway to kinetic friction, the energy
stored in the draw cables is released brutally, and the deck advances suddenly. Tens of launch
cycles compromise strand isotensioning, and pull adjustment is soon necessary. The pull in the
strands is equalised with single-strand jacks at the dead anchors so as not to interfere with
launch operations.

There are numerous possible schemes for applying the thrust force to the deck, and numerous
physical principles for load transfer. The use of friction between the deck and the anchor plates
of the draw cables, mobilised by hydraulic compression as in Figure 2.23 or by through bars as
in Figure 2.24, is limited to short PC bridges and minimal launch gradients, which require small
thrust forces. Higher thrust forces require mechanical transfer by means of thrust beams applied to
the rear end of the deck or crossing the box girder vertically.

g Horizontal thrust beams applied to the rear end of the deck provide centroidal thrust but
complicate web reinforcement and require anchor brackets on the launch abutment to lift
the draw cables to the centroidal deck level.

46

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


Bridge launching

Figure 2.23 Friction transfer of the thrust force by hydraulic compression. (Reproduced with permission
from ASCE)

Anchor plate of draw cable

Neoprene Contact surface

Jack

Deck

Return roller

g Vertical thrust beams and draw cables located beneath the deck apply a couple to the
deck (Figure 2.25). When the bridge is long and the thrust force is therefore high, the
couple may be excessive for young segments. Vertical thrust beams simplify reinforcement
at the construction joint, but require forming and sealing of two block-outs in every
segment.

In both cases, the launch jacks may be applied to the launch abutment or to the thrust beam. On
segment completion, the thrust beam is applied to the deck, and the draw cables are brought back

Figure 2.24 Friction transfer of the thrust force by means of prestressing bars. (Reproduced with
permission from ASCE)

Beam hangers

Prestressing bars

Thrust beam Draw cable

47

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


Bridge Launching

Figure 2.25 Vertical thrust beam crossing the deck. (Reproduced with permission from ASCE)

Removable thrust beam

Elastomeric pads

Draw cable

to their initial length. The weight of the cables complicates this operation beneath the deck, and
long bridges often justify the use of electrical trolleys for repositioning (Gillet and Jacquet, 1988).

Uphill launching requires deck anchoring during the return strokes of the launch jacks. The jacks
for strand jacking have two hydraulic clamps that avoid relief of the draw cable during reposition-
ing of the plunger. Releasing the load is less safe than pulling, as the front safety grips must be held
open. The jack may be gimbal-mounted for uniform load distribution among the strands and to
avoid angle breaks when launching along curves. The grip mechanisms are accessible at any time
(Rosignoli, 2013).

The capacity of the strand jacks ranges from 0.15 to 7.5 MN based on draw cables of 150 strands
of 18 mm diameter, seven-wire, die-compacted prestressing strand having a guaranteed breaking
load of 0.38 MN/strand. The stroke of the jacks varies between 250 and 500 mm. The load is held
mechanically when movement is stopped at any point of the stroke.

The launch of a steel girder is slower than the cycle time of the strand jack due to the time required
for clearing eld splices past rollers, rolling off temporary supports, coordination of personnel and
crew rotation. Launching a PC box girder is more regular but is slowed down by the ow rate of

48

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


Bridge launching

non-specialised hydraulic pumps (Rosignoli, 2013). In either case, strand jacking is slow compared
with other launch methods, typically in the 28 m/h range, with an average estimate of 4 m/h. The
launch speed of paired jacks can be synchronised irrespective of the size of the load.

Prestressing jacks are less expensive than the specialised systems for strand jacking but typically
have only one grip mechanism, and anchor plates are therefore necessary to lock the strands at
the end of each launch cycle. This solution is inexpensive but prevents backwards movement of
the deck in case of need. When two draw cables are used for launching, one cable provides the
longitudinal restraint during repositioning of the jack of the other cable.

In the heaviest applications, it is often preferable not to anchor the deck with the draw cables but
to use specic lock devices based on friction. Friction locks ensure higher safety when it is
necessary to replace some strands or to equalise the pull in the strands because of wedge slippage.
Friction locks are also used during casting of deck segments when thrust beams are used for
launching and the deck has a natural sliding direction. Because of the variability in the friction
coefcient of PTFE, the safety locks are designed disregarding friction at the launch bearings
when favourable for equilibrium. Friction anchoring at the abutment may be achieved in different
ways:

g The deck is lifted with jacks anchored to the abutment and equipped with oversized seals to
resist high shear forces. The jacks have tilt heads and knurled contact plates to mobilise a
high friction coefcient against the deck soft.
g The deck is lifted, the launch bearings are cleaned from grease, the Neoon pads used for
launching are replaced with elastomer pads or steel plates devoid of PTFE surface, and the
deck is released onto the launch bearings.
g The launch bearings are permanently placed on jacks, which are retracted on launch
completion to release the deck onto RC support blocks equipped with knurled contact
surfaces.

These lock systems are less effective in the nal launch phases, as their longitudinal shear capacity
depends on the continuous beam support reaction at the abutment. The latter remains almost
constant during launching but decreases in the nal launch phases, when the weight of the deck,
and with it the sliding force to be controlled, is at its highest.

Some PC box girders have been anchored with transverse jacks pressing against the sides of the
bottom slab. The shear capacity of this type of restraint does not depend on the support reaction
and is therefore stable at the end of launching. The effectiveness of the restraint depends on the
contact pressure, and the jacks must therefore be equipped with mechanical ring nuts to prevent
loss of restraint in case of collapse of the hydraulic system.

Slopes steeper than 23% may suggest downhill launching. The braking force is smaller than the
thrust force required for uphill launching, as friction works in favour of equilibrium. If the braking
devices are designed to pull the deck uphill (this facilitates the release of temporary lock pins and
the correction of an accidental excessive advance), the load advantage is lost. The exural stiffness
of the piers opposes thermal deck contractions during the launch stoppages for construction of
new segments. When the braking devices are designed considering this additional load, the launch
bearings may be placed on thick elastomeric pads to diminish the combined longitudinal stiffness
of the piers.

49

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


Bridge Launching

In downhill launching with 23% gradient, the deck must be pushed to initiate movement, and
then held back when the breakaway friction has been overcome. Two xed braking cables may
be anchored beneath the deck between the casting cell and the launch abutment, and one central
draw cable is anchored from a vertical thrust beam to the abutment. The thrust beam crosses the
deck and is connected to a bottom cross-beam that anchors three strand jacks. Two front jacks are
used for braking, the rear central jack is used for the initial push, and a common hydraulic system
controls jack synchronisation (Bennett and Taylor, 2002). Lighter decks may be launched with
two strand jacks working nose to nose on the same xed cable; the braking jack is applied to the
front face of the thrust beam, and the thrust jack is applied to the rear face. Two xed cables, two
pairs of jacks and two thrust beams are used for redundant operations. This scheme may be
reversed by anchoring bars or cables to the thrust beam, by releasing the bars from the casting cell
for braking, and by pulling the bars from the launch abutment for pushing the deck forward.

The main limitation of draw bars and cables is their capability of working only in tension. Draw
bars and cables may be replaced with a perforated steel plate that runs under the deck from the
thrust beam to the launch abutment. Two double-acting launch cylinders anchored to the abut-
ment push the plate forward by means of a saddle supporting a launch pin that crosses the plate.
When the launch cylinders reach the end of stroke, a lock pin is inserted into the plate to restrain
the deck, and the launch pin is extracted to reposition the saddle for a new stroke. Hydraulic actua-
tors on the lock and launch pins and electronic switches that conrm pin insertion can be used to
automate the launch cycle.

The perforated plate has rectangular cross-section, and its axial stiffness provides some extent of
bidirectional restraint. When a more reliable bidirectional restraint is needed, the perforated plate
is anchored to the casting cell and the launch abutment, and the launch saddle is anchored to the
vertical thrust beam to push or brake the deck as necessary.

Tow systems present some limitations when the bridge is curved in plan, as the thrust force is
applied along the chord between the thrust beam and the launch abutment. This produces angle
breaks in the draw cables at their entry into the launch jacks and the dead anchorages, which
require the use of multiple cable deviators (Figure 2.26).

Figure 2.26 Deviators and lateral guides are necessary for the draw-cable launch of curved decks.
(Reproduced with permission from ASCE)

2nd phase casting cell Curing support

1st phase casting cell Abutment

Draw cable

Lateral guide Deviator

Continuous foundation Launch jack

50

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


Bridge launching

Figure 2.27 Rear 3.2 MN thrust beams with prestressing jacks

The draw cables are rarely designed to slide within the deviators, and the launch jacks are therefore
applied to the thrust beams (Figure 2.27). The deviators are progressively removed during launch-
ing to permit the launch jacks to pass through. Applying the thrust force along the chord also
tends to cause drifting in the rear deck end, which requires lateral guides in the casting cell and
the curing area. Deck launching along an inclined cone causes additional drift forces due to
cross-fall distortion at the deck soft. Finally, drawbars and draw cables can rarely be located
alongside the deck because of the interference with the deck surface at the outer side of the curve.

In spite of the attention required in particular cases (steep gradient, curved alignment), operative
slowness, the higher labour demand and the large geometry adjustment of the casting cell to avoid
conicts with the draw cables, tow systems are often the least expensive and most efcient solution
for the launch of short PC decks and medium-length steel girders. They are based on simple and
reusable equipment, and are easy to assemble and dismantle.

2.8.1.2 Rear thrust systems


As an alternative to the tow systems, long-stroke double-acting hydraulic cylinders can be applied
to the rear end of the deck to transfer the thrust force by acting against reaction beams. The foun-
dation beams of the casting cell are used as reaction beams for the launch of PC box girders, and
two rear thrust systems are therefore used to move the deck without load eccentricity and for alter-
nate repositioning. A central pair of thrust cylinders may be sufcient for the horizontal launch of
single spans (Figure 2.28).

Rear thrust requires anchoring the thrust cylinders at many locations along the reaction beams.
Concrete reaction beams are engaged with through pins or by means of hydraulic compression
(Figure 2.29). The clamping force is about three times the peak thrust force applied to the deck

51

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


Bridge Launching

Figure 2.28 Rear thrust cylinders applied to the central foundation beam

Figure 2.29 Rear thrust cylinder with hydraulic clamps. (Reproduced with permission from ASCE)

Deck Thrust cylinder

Reaction beam

Clamping jacks

Lock

Launch

Repositioning

52

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


Bridge launching

Figure 2.30 Rear thrust cylinders and continuous rack. (Reproduced with permission from ASCE)

Deck

Reaction beam

based on a friction coefcient Cf,L h 0.6 and a resistance factor f = 0.5. The modular clamp of the
launcher may be lengthened with additional modules to provide larger thrust forces. The launch-
ers are anchored to the rear end of the deck for automatic repositioning. Backward deck pulling or
braking during downhill launching require powerful anchor systems to the deck. Long thrust
cylinders with an effective stroke of about 1 m are used to diminish the number of launch cycles.

A steel rack may be embedded in the top surface of the reaction beam to restrain a launch tooth,
and this provides a safe and less expensive mechanical load transfer (Figure 2.30). The small thrust
reaction of steel girders can be resisted with steel reaction beams anchored to the launch abutment.
Double-acting thrust cylinders bolted to the rear eld splice of the girders are anchored to the
reaction beams with through pins. One thrust cylinder is used on each girder for redundant launch
operations and to avoid overloading of lateral bracing due to differential movements of the steel
girders (Figure 2.31). The use of cylinders with eye-bar plungers allows backward pulling if the
reaction beams are anchored at both ends.

The rear thrust systems are typically used for the horizontal launch of steel girders, prestressed
composite box girders, and medium-weight PC decks. The higher thrust forces required for heavy
PC decks or for uphill launching cause several difculties.

g The deck can rarely be pulled backwards, as this would require many anchor bars between
the thrust cylinders and the rear end of the deck at every construction joint. Pulling a steel
girder backwards is much easier, as the girder ends with a eld splice, and when the eld
splice are bolted numerous holes are available in the web and the bottom ange to anchor
the thrust cylinders.
g The thrust cylinders slow down setting-up of the casting cell for the next segment, and
interfere with the distribution of reinforcement and launch tendons in the segments.
g The casting cell must be located close to the abutment to transfer the thrust reaction to its
foundation. Ensuring minimal angle breaks at the rear end of the deck for match-casting of

53

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


Bridge Launching

Figure 2.31 Redundant thrust cylinders for a girdersubstringer system. (Photo: LaViolette)

the next segment often requires a temporary pier in the end span of the bridge, and adds to
the costs.
g The thrust reactions apply negative couples to the reaction beams, and induce axial
tension. Axial tension suggests the use of longitudinal prestressing in the reaction beams
to minimise cracking and to ensure full exural stiffness. Migration of the thrust couples
along the reaction beams suggests the use of pile foundations to prevent settlement of
the casting cell, which would jeopardise deck geometry and the vertical alignment of the
entire casting yard. Pile foundations and prestressed reaction beams increase the cost of
the casting facility.

The rear thrust systems offer several advantages over tow systems. Launching is faster, and this is a
major advantage because 23 workers are necessary at every pier cap during launching, and saving
on labour costs soon breaks even with the cost of special launch equipment and reaction beams.
The launch of curved decks is also easier, as the thrust force is applied along the local tangent, and
only 12 workers are needed to operate the thrust devices. Finally, bidirectional anchor racks and
friction clamps allow hydraulic braking of the deck during downhill launching.

The cost of the rear thrust systems increases rapidly with increasing deck weight and launch
gradient, and these systems soon become uneconomical. In long PC bridges, rear thrust systems
have been used to overcome breakaway friction at the beginning of launch in combination with
tow systems designed for kinetic friction. In most cases, however, heavy PC box girders suggest
the use of launch devices that transfer the thrust force by friction.

2.8.2 Eberspacher launchers and derived systems


In the heaviest bridges, the thrust force is applied by friction using one or more synchronised pairs
of electro-hydraulic launchers that support the deck under the webs. In the simplest version, an
Eberspacher launcher includes a vertical jack pushed along a lubricated surface by a longitudinal
double-acting cylinder. The thrust cylinder is anchored to a rear reaction block that supports the
deck in the pauses between two subsequent launches (Figure 2.32).

54

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


Bridge launching

Figure 2.32 Basic configuration of an Eberspacher launcher

In the most advanced version, a friction launcher is a monolithic device that includes a sliding
sledge containing multiple support jacks interconnected hydraulically to generate a spherical
hinge. The tilt heads of the jacks are welded to a knurled contact plate that supports the deck
under the webs with a friction coefcient Cf,L h 0.6. The contact plate is machined to transfer
support reaction and thrust force without stress concentration and damage to the deck surface;
this is usually achieved by using tight inclined cuts to transfer the support reaction and thrust force
by local bending (Rosignoli, 1998a).The bottom surface of the sledge lodges a dimpled PTFE plate
that slides longitudinally along a polished stainless-steel plate integral with the basement of the
launcher. One or two long-stroke double-acting cylinders pinned to a rear reaction block provide
the thrust force to the sliding sledge (Figure 2.33) (Rosignoli, 1998a).

In most applications, two synchronised launchers are installed on top of the abutment. All types of
Eberspacher launcher include two interfaces for transfer of the deck support reaction: a top high-
friction interface between the deck and the launcher, and a bottom low-friction interface between
the launcher and the abutment. Longitudinal rails may be applied to the launcher to guide the
movement of the sliding sledge and to provide transverse restraint to the deck by friction. If the
deck has tight plan curvature, the lateral launch guides are applied to the abutment to minimise
the lateral load applied to the launcher by vector decomposition of the thrust force.

Pairs of friction launchers have reached 26 MN of lifting capacity and 9.2 MN of factored thrust
capacity. A resistance factor f = 0.5 is typically used for the design of launch operations to ensure
reliable control of deck movements, and the ratio of the lifting capacity to the factored thrust capacity
is therefore around three. Figure 2.34 illustrates the working cycle of an Eberspacher launcher.

55

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


Bridge Launching

Figure 2.33 Central Eberspacher launcher for fish-belly cross-sections. (Reproduced with permission
from ASCE)

1 The support jacks lift the deck 510 mm from the reaction block of the launcher. The
support reaction of the deck is thus transferred from the reaction block to the sliding sledge
of the launcher.
2 The thrust cylinders push the sledge forward along the steelPTFE sliding surface of the
launcher. The thrust force is transferred to the deck by friction, taking advantage of the
deck support reaction on the sledge and the high friction coefcient of the knurled contact
plate placed on top of the support jacks.
3 The effective stroke of the thrust cylinders varies between 0.25 and 1.00 m. When the end of
stroke is reached, the support jacks are retracted to lower the deck onto the reaction block.
4 The thrust cylinders pull the sledge backwards beneath the deck to the initial position to
start this cycle again.

The tilt head of the support jacks copes with exural rotations in the deck and the launch gradient.
Linear position sensors applied to thrust cylinders and support jacks provide operation feedback
to a programmable logic controller (PLC) for launch cycle automation. Pressure sensors control
the hydraulic systems to stop the launch sequence if pre-set limit pressures are reached.

The ratio of the thrust force Ft necessary to move the deck to the factored deck support reaction
fRv,L on the launcher must be smaller than the friction coefcient Cf,L between the knurled
contact plate of the launcher and the bottom surface of the deck.
Ft
Cf,L (2.1)
fRv,L

56

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


Bridge launching

Figure 2.34 Working cycle of a friction launcher. (Reproduced with permission from ASCE)

Deck

LIFTING

Abutment

THRUST

LOWERING

RETURN

As both Ft and Rv,L depend linearly on the weight of the deck cross-section, the latter does not
inuence the launchability criterion expressed by Equation 2.1, which depends only on the fric-
tional resistance of the launch bearings and the casting cell, on the average launch gradient, and
on the tributary deck length on the launcher (i.e. on the position of the launcher with respect to the

57

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


Bridge Launching

adjacent supports). In very long PC bridges, high thrust forces require a signicant distance
between the launch abutment and the front support of the curing area. This lengthens the casting
yard, restrains the distribution of the curing supports between the launch abutment and the casting
cell, and inuences the sequence of application of launch prestressing. When the rst pier is not
very tall, therefore, synchronised pairs of friction launchers may be used at the abutment and
at the rst pier to increase the tributary deck length on the launchers and the total load Rv,L with it.

The support jacks lift the deck by 510 mm from the reaction blocks during launching. The
upward deection is considered in the analysis of the launch stresses in the deck, may or may not
be considered in the evaluation of Rv,L , and is set as an end-of-stroke process variable in the PLC
software program. Extra lifting can increase the support reaction if necessary, if compatible with
the level of curing and prestressing of the deck.

The support jacks are generally oversized to lift the deck, even when misaligned launch bearings or
thermal gradients increase Rv,L . The thrust cylinders are designed for the peak thrust force, taking
into consideration the extraction friction of the segment from the casting cell, the breakaway
friction of the launch bearings and the launchers themselves, and the launch gradient during the
individual launch operations.

In most cases the friction launchers are used only at the abutment, and the thrust force denes the
tributary deck length on the launchers. Launch prestressing is fully applied before the segments
reach the end span of the bridge. The launch tendons cross 23 deck segments, and the casting cell
is therefore far from the abutment. A curing area with additional temporary supports is inserted
between the casting cell and the launch abutment to complete curing of segments and the
application of launch prestressing. This geometry of the casting yard requires specic operations
in the initial and nal phases of launching.

At the beginning of launch, the curing area separates the casting cell from the launch abutment.
The need for structural continuity between the launch nose and the deck is typically solved by
match-casting (i.e. by assembling the launch nose at the front bulkhead of the casting cell). The
launch nose and deck must therefore be towed onto the launchers until a support reaction
sufcient to continue the launch by friction is obtained. For this purpose, draw cables are used
to pull the nose with the thrust cylinders of the launcher (Figure 2.35).

In the nal phases of launch, Rv,L decreases until the end support reaction of the continuous beam
is reached. This end support reaction is often insufcient to satisfy the launchability criterion
expressed by Equation 2.1 under the highest thrust force demand (Rosignoli, 1998b). When the
support reaction at the launch abutment is insufcient for several launches, a second pair of
Eberspacher launchers is applied to a front cable-stayed pier. When the load is insufcient only
in the last 1030 m of launch, a steel nose is applied to the rear end of the deck to pull draw cables
with the thrust cylinders of the launcher (Figure 2.36). The draw cables are designed to integrate
the thrust force provided by the launcher through friction. Draw cables tensioned by prestressing
jacks may also be used to pull the deck at the target abutment (Figure 2.37).

The use of synchronised friction launchers controlled by PLC networks may generate enormous
coherent thrust forces. Such thrust capacity may be indispensable for uphill launching of long PC
bridges on steep gradients. In addition to redundancy and accurate distribution of the launch
forces, synchronisation multiplies the number of reaction points and avoids concentrating such

58

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


Bridge launching

Figure 2.35 Initial launch phases of the Aronde Bridge

forces on just one point. The number of devices to pilot, the small tolerances of launch operations,
the massive feedback provided by multiple launchers, and the longitudinal and vertical exibility
of the reaction points make such launch systems ungovernable manually. Redundant PLC net-
works ensure the reliable control of synchronisation, well beyond the possibilities of human

Figure 2.36 Rear nose and integrative draw cables for the Serio River Bridge. (Reproduced with
permission from ASCE)

59

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


Bridge Launching

Figure 2.37 Final front towing of the Palizzi Bridge. (Reproduced with permission from ASCE)

operators. The PLC program informs the operator on the launch parameters and allows correction
of parameters accessible to the operator, and emergency interruption of the launch sequence.

Compared with tow and rear thrust systems, the use of friction launchers offers many advantages.

g Failure of a non-redundant tow or rear thrust system would leave the deck unrestrained on
a low-friction inclined plane. Design for redundancy and the frequent use of higher safety
factors result in oversized equipment and slow operations. The friction launchers ensure
absolute intrinsic safety, as the worst consequence of mechanical or hydraulic failure would
be launch stoppage and descent of the deck onto the reaction blocks of the launchers,
which are designed as frictional restraints. The friction launchers can be designed for the
launch loads, and can be overloaded without excessive concern if necessary. The friction
launchers are also able to pull the deck backwards if required, and are compatible with
launching uphill, downhill and on 13% downhill gradients that involve thrust reversal
when passing from breakaway to kinetic friction.
g Friction launchers piloted by the PLC allow real-time recording of the launch parameters.
By monitoring Rv,L it is possible to verify the accuracy of deck geometry and to correct
errors to prevent accumulation. By relating Rv,L to the deck lifting it is possible to evaluate
the elastic modulus of the young concrete. Setting an upper limit for the thrust force avoids
pier overloading in the case of seizing or upside-down insertion of a Neoon pad.
g The launch cycle can be automated with linear displacement transducers, contact switches
and pressure valves that provide feedback to the PLC that drives the control valves on the
feeding and return lines of thrust cylinders and support jacks. PLC control facilitates the

60

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


Bridge launching

operations, which can be supervised by one technician only. It also increases the launch
speed, which can amply exceed 10 m/h (i.e. less than 3 hours to extract the segment from
the casting cell). Thrust cylinders and support jacks may be driven with a higher ow rate
during the return strokes to shorten the cycle time, and the support jacks may be slowed
down during the nal approach to the deck to avoid surface damage.
g The faster launch operations and the lack of need to reposition the draw cables and
anchoring the deck during these operations diminish the labour demand of launching.
g The casting cell is immediately ready for the construction of the next segment. When a
portal crane is used to transfer the prefabricated cage from the rebar jig to the casting cell,
deck launching, form cleaning, positioning of the Neoon pads on the extraction rails, and
cage insertion can often be completed in the same day.

A disadvantage of friction launchers is that, when fully extended, the thrust cylinders are 34 times
longer than their effective launch stroke. Long effective strokes are necessary to shorten the launch
duration, reaction block and sliding sledge add length to the basement, and the total length
increases so much that the abutment wall must often be 34 m thick or more. Another disadvan-
tage is the complexity of the operations for removing such heavy devices from the abutment at the
end of launching.

The original Eberspacher launchers have evolved into several different schemes. The Millau
Viaduct in France was displaced using electro-hydraulic launchers placed on each temporary and
permanent pier, the process being controlled by a PLC network. Each pier was equipped with a
power-pack unit and local PLC control panel, and a remote PLC master control panel was used
to monitor the local panels for synchronised operations. Three launch modes were provided:
manual for adjustment and small local corrections, semi-automatic for step-by-step movements,
and automatic for the full launch cycle. Each launcher included a base frame supported on
long-stroke vertical cylinders that provided vertical adjustment and controlled rotation in the
longitudinal and transverse planes. The base frame was equipped with a low-friction surface for
the movement of a bottom lifting wedge, which was driven by double-acting push cylinders
anchored to the base frame. The top inclined surface of the lifting wedge was equipped with a
second low-friction surface for the movement of a top launch counter-wedge along the lifting
wedge, which was driven by double-acting pull cylinders also anchored to the base frame. The top
wedge supported the deck on its top horizontal surface. The working cycle of the launcher
included pushing the lifting wedge backward to lift the launch wedge and support the deck, pulling
the launch wedge and deck forward along the lifting wedge, pulling the lifting wedge forward to
lower the launch wedge and release the deck, and pushing the launch wedge backward for a new
cycle. The system provided 600 mm effective stroke with 4-minute cycle time.

The launcher used for the Olifants River Bridge in South Africa included double-acting thrust
cylinders driving a mobile thrust beam that rolled beneath the deck. Transverse jacks on the
thrust beam were used to clamp the bottom slab of the PC box girder to enable frictional load
transfer through the outer surface of the webs. Conventional launch bearings were used to support
the deck at the launch abutment. The working cycle of the launcher included deck clamping,
thrust, release and return of the thrust beam to the initial position. These types of launcher do
not depend on the support reaction, simplify the distribution of the temporary supports in the
curing area, and provide full load capacity also in the nal phases of launch. These launchers,
however, require deck anchoring during repositioning of the thrust beam when launching along
inclined planes.

61

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


Bridge Launching

Figure 2.38 Friction launcher with flange clamps and wedge brakes. (Photo: Somerset/KWH)

Similar concepts have been used for the design of friction launchers for steel girders. Because of
the light weight of the girder, the friction launchers are often equipped with hydraulic ange
clamps. High contact pressure diminishes the transfer surface of the thrust force, and the eld
splices in the bottom ange of the girder are designed to avoid interference with the clamps.
Friction launchers equipped with ange clamps (Chemerinski et al., 1996; Popov and
Seliverstov, 1998) are often combined with wedge brakes (Gale, 2011). The integrated launcher
shown in Figure 2.38 includes, from right to left, an articulated launch bearing based on two
Hillman rollers on an equalising beam, the wedge brake, the thrust cylinder, and the hydraulic
ange clamp.

The clamps may be suspended from the bottom anges of the steel girder during the return stroke
of the thrust cylinders, or may be supported on return carriages. The launch cycle with wedge
brakes includes the application of ange clamps, release of wedge brakes, extension of thrust cylin-
ders, application of wedge brakes, release of ange clamps and re-entry of thrust cylinders (Gale,
2011). Because of the critical implications of friction braking on the stability of equilibrium, the
launch systems are load tested prior to their use on steep gradient applications.

Synchronised thrust cylinders are used on every girder in order not to overload the lateral bracing
between the girders with differential launch movements. In the absence of wedge brakes, one
thrust cylinder locks the steel frame during repositioning of the other cylinders in order to avoid
uncontrolled movements. Thrust cylinders with a 2 m effective stroke and high ow rate power-
pack units have been used to diminish the number of launch cycles. The thrust cylinders may
be used to rotate vertical arms pinned at the bottom so as to amplify the longitudinal displacement
of the hydraulic clamps; these launchers have been used for steel girders weighing more than
10 000 tonnes (Zhuravov et al., 1996).

2.9. Launch bearings and lateral guides


The launch of PC box girders, prestressed composite box girders and steel I- and U-girders
requires different types of launch bearings, which are designed for the different weight and exi-
bility of the deck and the different geometry of the webs. Because of the peculiarities of design, the

62

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


Bridge launching

launch bearings for PC bridges and the prestressed composite box girders are discussed in
Chapter 3, and those for steel girders are discussed in Chapter 4.

The use of lateral guides is also necessary during launching, to maintain the correct alignment of
the deck and to resist lateral forces due to wind and the seismic design demand during construc-
tion. The mass of long PC decks requires strong lateral guides, especially when the deck is
launched along a curve. Strong lateral guides are also necessary to resist the seismic design event,
while the effects of transverse wind are generally modest. The lateral guides for light steel girders
are often designed for lateral wind.

Although all the piers of a rectilinear PC bridge are equipped with lateral guides, deck alignment is
corrected at two points only: at the front end of the casting cell (to correct the position of the
construction joint with regard to the next segment) and at the lead pier reached by the PC deck
(to adjust the direction of the launching nose at landing at the next pier). The system is isostatic and
the guide forces are minimal. Curved PC decks are guided at multiple locations, and vector
decomposition of the thrust force increases the guide forces. Large guide forces also result from the
soft cross-fall of solid and voided slabs (De Clercq and De Ridder, 2003). Light steel girders are
also guided at every pier to resist lateral wind and to keep the webs aligned with the launch bearings.

The lateral guides are designed for the loads necessary to move the deck laterally by acting at the
most appropriate points during each launch stage. To account for the need for adjustment of
the deck alignment, some design criteria for PC bridges require that the calculated lateral loads
be increased by 1% of the support reaction at the pier. At least 3% of the maximum support
reaction plus the lateral reaction to lateral wind should be used when launching a steel girder
(Rosignoli, 2000a).

The guide forces are determined based on experience and vector decomposition of the thrust
force in bridges curved in plan. Launch guides designed to restrain the deck during the seismic
design event for the construction stages are often oversized for the launch requirements. When the
lateral guide acts against an inclined web surface, the guide force is perpendicular to the web
surface and its vertical component reduces the support reaction produced by the adjacent launch
bearing.

Lateral deck equilibrium requires at least two guide points. The rear guide is applied to the
front end of the casting cell, and the front guide is applied to the lead pier. The guide force
diminishes as the distance between the two guide points increases. Using multiple guide points
is not recommended for a PC deck as the response of hyperstatic guide systems depends on the
transverse stiffness of the piers and is hardly predictable. Lateral guides are, therefore, applied
to all of the piers for rapid intervention in case of need, but only two guides are kept engaged
by inserting the Neoon plates. Multiple guide points are typically used for the launch of steel
girders because of the need to keep the support reactions aligned with the webs. Additional guide
points are necessary in PC decks with tight plan radius in order to minimise horizontal deck bend-
ing and the lateral displacements at the launch bearings. Guide points at all piers are used for the
launch of curved PC slabs with soft cross-fall (De Clercq and De Ridder, 2003).

When horizontal or uphill launching takes place along a plan curve, the outer guides are subject to
radial forces resulting from the vector decomposition of the thrust force. When launching down-
hill, the deck is braked, the direction of the guide forces reverses, and the guides are loaded on the

63

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


Bridge Launching

inner side of the curve. When the gradient of downhill launching is intermediate between break-
away and kinetic friction, the deck is pushed until detachment and then braked, and the direction
of the radial guide forces inverts during launching.

The design of the launch guides for lateral wind is simpler. The design standards specify the trans-
verse wind load, which is applied to the deck with the least favourable distribution to determine
the lateral demand on the guides. Transverse wind typically governs the design of lateral guides for
steel girders, while its effects are smaller when launching heavy PC decks. Alignment forces are
added to the wind loads determined for the maximum wind speed allowed during launching.
Exceptional wind conditions may require interruption of the launch; in this case, the load on the
lateral guides is assessed without alignment forces.

The seismic design of the lateral guides is more complex. Most design standards specify the seismic
design demand in terms of peak ground acceleration (PGA) and spectral modication factors that
depend on the importance of the structure and on the classication of the site. The PGA for the
seismic design of the bridge reects a seismic event with a given return period. In many cases the
return period is 475 years and the seismic design event has a probability of exceedance ranging
between 10% and 19% for a design life of the structure of 50 and 100 years, respectively.

The seismic demand during construction is determined based on the probability of exceedance
during construction and on the construction duration. Eurocode 8 (BSI, 2005b) species the
return period to be used to determine the PGA during construction. If one assumes a duration
of construction of 1 year and accepts a 5% probability of exceedance, the return period of the
design event during construction is 20 years, and the PGA during construction varies between
24% and 39% of the PGA for the structure in service. In regions of high seismicity, these levels
of seismic demand may be signicant, and have different effects in the longitudinal direction and
the transverse direction.

In the longitudinal direction, the deck is supported on low-friction launch bearings. Neglecting
friction, the longitudinal bridge response must be resisted with specic restraints. The friction
launchers also act as frictional restraints, but when the deck is launched with tow systems or rear
thrust cylinders, special seismic restraints are necessary to lock the deck, even in the case of hori-
zontal launching. The most logical location for the seismic restraints is at the launch abutment,
although a stiff restraint shortens the longitudinal period and increases the spectral demand.
Seismic restraints may also be lodged at some piers to take advantage of pier exibility, although
deck locking is more complex and may result in pier overdesign in the case of isolated bridges.

In the transverse direction, on launch completion the deck is connected to the pier caps by means
of transverse shear keys or seismic isolation systems. In the rst case, piers designed for the trans-
verse response to the full PGA in service conditions offer wide margins during construction. In the
case of an isolated bridge, the piers are designed for the transverse demand transferred by the
isolation system and are not protected during construction. However, the PGA during construc-
tion is smaller than the PGA in service, and this is often enough to avoid pier overdesign. In both
cases, the lateral guides are designed according to the capacity design principles (Calvi et al., 1996)
with adequate protection factors.

The launch bearings used for deck rotation are conceptually similar to those used for frontal
launching. Their design is complicated by the need to ensure longitudinal and transverse

64

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


Bridge launching

stability of the deck during rotation. If the deck has high torsional stiffness, transverse stability
may be assigned to the rear balancing frame. The frame is rigidly connected to the deck,
includes two braced columns that resist the torsional moment, and is located so as to obtain a
high support reaction. Ribbed slabs or composite decks have low torsional stiffness, and trans-
verse stability is achieved by restraining the deck torsionally at the pivot pier. Three approaches
are possible.

g The pivot pier provides longitudinal and transverse stability, like in a swing mobile bridge.
This scheme is rarely used in PC decks because of the load imbalance deriving from the
short lever arms at the pivot pier. This scheme may be the only solution when the area
surrounding the deck is inaccessible (a pier inside a river or between roads or railway
tracks). In this case, the deck should be as symmetrical as possible.
g The pivot pier provides transverse stability, and the rear balancing frame provides
longitudinal stability. This scheme applies the torsional constraint through the most
loaded support, and the balancing frame may be a simple strut. The support triangle thus
obtained avoids indeterminate effects during rotation. If the torsional stiffness of the deck
is insufcient, the rear balancing frame may also provide some degree of torsional
constraint.
g The rear balancing frame provides longitudinal and transverse stability. A compact rotating
bearing may be used at the pivot pier to create a support triangle. This scheme requires
high torsional stiffness in the deck, simplies the pivot bearing, and permits pier design for
the nal deck geometry, leaving rotation out of consideration. This scheme is the most
efcient and economical but requires a rigid pier diaphragm that transfers the whole
support reaction to the deck webs during rotation.

In the rst two schemes, the rotating support at the pivot pier includes Neoon plates that slide
along polished stainless-steel circular crowns. If the loads are low, the steel crowns may be
embedded in the foundation of the pivot pier. Higher loads suggest embedding the steel crowns
in the deck and bolting Neoon circular crowns to the foundation. A guide is also necessary to
keep the deck centred with respect to the pivot pier. Two concentric steel pipes separated by
Neoon rings can be embedded in the pivot pier to reduce friction and to avoid relative move-
ments. The rotating support should be designed and built carefully, as replacing defective elements
is very difcult. The deck segment over the rotating support may be cast with the nal alignment
and rotated to the deck casting position to check proper functioning without the weight of the
deck.

The compact rotating bearings for the third scheme include a Neoon plate placed over a stainless-
steel plate, both crossed by a removable vertical pin. Similar bearings were used in the Danube
Bridge in Austria and in the La Fleche Bridge in France. Hydraulic jacks with a rotating plunger
were used for the Fontenelle Bridge and the Trith Saint-Leger Bridge in France. Hydraulic
jacks ensure several functions at the same time (deck lifting, rotation, centring, lowering over the
permanent bearings, dampening of vibration and reduction of friction) but require careful
construction for full-load rotation.

The balancing frame includes a pair of braced columns that slide over a circular steel rail by means
of lubricated steel sledges. Higher loads require the use of stainless-steel saddles over Neoon
plates arranged along a concrete runway beam. This type of support is more exible, and the
launch movement more regular.

65

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


Bridge Launching

The tangential force necessary to rotate the deck is applied as far as possible from the pivot pier. In
the second and third support scheme, it is applied to the balancing frame. Low friction at the pivot
pier is necessary to avoid secondary stresses and deck vibration. The small launch force can be
obtained by means of light self-clamping hydraulic cylinders, the action of which is more uniform
than that of pulling bars or strands.

2.10. Launch and lock forces


The longitudinal force necessary to move the deck or to prevent its uncontrolled sliding depends
on the average gradient of the launch surface and on the friction resistances that oppose
movement.

The effects of the launch gradient are easy to evaluate. Setting Q as the weight of the deck section
being launched, the longitudinal gradient, tg a, of the launch surface at the centre of mass of the
deck denes the average launch gradient. The horizontal force Fg to be applied to the deck section
to launch it or to lock it along a frictionless launch surface is
Fg = Qtga (2.2)
The average launch gradient is calculated for the individual launch phases. When the vertical
curvature radius is small, the average launch gradient can be very different from the nal gradient
between the abutments. The local gradient of the Tiziano Bridge in Italy (see Figure 2.6) is +5.3%
at the launch abutment and 5.3% at the opposite abutment, and the average gradient is therefore
zero (Rosignoli and Rosignoli, 2007). During launching, the average launch gradient was +7.2%
at the beginning of the rst 50 m launch (with the deck entirely behind the abutment), +5.8% at
the beginning of the second launch (100 m deck), +4.4% at the beginning of the third launch
(150 m deck) and +3.1% at the beginning of the nal launch. The cable-stayed Palizzi Bridge
(Rosignoli, 1998d; Martinez Y Cabrera and Rosignoli, 2001) was in a similar situation (see
Figure 2.17). Calculating the average launch gradient for the individual launch phases is indispen-
sable for the safe design of the launch devices and for prevention of uncontrolled deck sliding.

A PC deck is launched by sliding along different support surfaces, each of which has its own fric-
tion coefcient. The launch bearings at the piers typically provide most of the frictional resistance
during launching. When the launch surface is horizontal, the peak longitudinal force F 0i necessary
to produce sliding at the launch bearing i depends on the local support reaction Rv,i and on the
breakaway friction coefcient C 0f,i of the launch bearing:

F 0i = C 0f,iRv,i (2.3)

PC decks, prestressed composite box girders, and the heaviest steel girders are launched on steel
Teon contacts. The behaviour of PTFE on polished stainless-steel surfaces is rather variable. The
friction coefcient diminishes from the breakaway friction C 0f,i to the kinetic friction Cf,i during
launching, and both these coefcients depend on several factors:

g Flatness and mirror polishing of the stainless-steel surface diminish both friction
coefcients in comparison with an unpolished, irregular surface.
g Dirt, poor lubrication and excessive wear of the PTFE pads increase friction substantially.
The friction coefcient for pure non-lubricated PTFE on stainless steel may be taken as
twice the value for lubricated contacts (BSI, 1983).
g Both friction coefcients decrease when the average contact pressure increases.

66

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


Bridge launching

Recommended friction coefcients for bridge bearings comprising pure continuously


lubricated PTFE sliding on polished stainless steel are 8% for a bearing pressure of 5 MPa,
6% for 10 MPa, 4% for 20 MPa and 3% for 30 MPa and above (BSI, 1983). The use of
small sliding surfaces for launching diminishes the launch force, facilitates insertion of the
Neoon pads under the deck, and keeps a lateral clearance from the deck edges to minimise
the risk of spalling of the unreinforced corner.
g An increase in temperature produces effects similar to an increase in the bearing pressure.

Although the friction coefcient diminishes with the progress of launching, the breakaway condi-
tion governs the pier design. A breakaway friction coefcient C0f,i = 0.05 applied to the peak
support reaction is adequate in most cases. Friction is neglected when determining braking or
locking forces, as kinetic friction may be very low (AASHTO, 2014).

The elastic shortening of the deck under the axial compression imparted by the thrust systems
staggers the breakaway of the launch bearings longitudinally. The bearings close to the launch
abutment slide rst and, when the front-most bearings start to slide, the friction coefcient at the
rear bearings is already tending to the kinetic friction. A contemporaneity factor fi , 1 may there-
fore be applied to the breakaway force at the rear bearings for the analysis of the total launch
force. This reduction is inuential only in long bridges.

The casting cell typically includes two xed extraction rails and a central form table that is lowered
before launching. The extraction rails are equipped with low-friction sliding surfaces. The friction
resistance Fcc opposed by the extraction rails is higher than the frictional resistance of the launch
bearings because of the low bearing pressure and the dirt and wear caused by grout spillage. The
difference between breakaway and kinetic friction is particularly marked in the casting cell. The
breakaway friction coefcient may be taken as 10% for Neoon pads and bakelised plates, and
15% or higher for greased steel plates. The extraction friction is applied only to the weight of the
segment, and is therefore more inuential in short bridges.

The frictional resistance of the curing supports is also higher than that of the pier bearings because
of the low support reaction. The friction resistance of lateral guides based on Neoon pads or roll-
ers is often negligible, even in the case of launching along curves, provided that the launch surfaces
are horizontal in the cross-section plane. Filled PTFE and metallic sliding materials frequently
used for the launch of steel girders may increase the guide friction substantially, but a steel girder
is typically lighter than a PC deck or a prestressed composite box girder.

The total frictional resistance Ff does not increase linearly with the number n of launch bearings
progressively engaged by the deck, and follows a law of the type


n
Ff = Fcc + fi Fi0 (2.4)
1

A load factor g . 1 is applied to the frictional resistance to allow for inadequate lubrication, wear of
the sliding surfaces, and the tendency of PTFE to seizing to the launch bearings during the launch
stoppages for casting of new segments. As few uncertainties affect the average launch gradient, Fg is
often used without load factors, and the design value of the thrust force is therefore

Ft,d = Fg + gFf (2.5)

67

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


Bridge Launching

During launching it is necessary to apply a system of forces to the deck to produce or to control its
movement. During the pauses of the launch, it may be necessary to apply a different system of
forces to avoid uncontrolled sliding. As no resisting friction is taken into account when designing
lock or brake devices, a load factor g 1.3 is applied to the effects of the average launch gradient
to determine the design lock/brake force Flb,d .

Flb,d = gFg (2.6)

In several highway bridges, the average launch gradient and the average friction coefcient are
similar. When Fg and Ff have the same sign (uphill launching), the sign reversal of Ff during the
launch stops may cause uncontrolled backward sliding of the deck when tow systems and rear thrust
cylinders are disengaged for repositioning. When Fg and Ff have opposite signs (downhill launch-
ing), prudence must also be exerted when determining the expected range of Ff . After breakaway,
new well-lubricated Neoon pads may have a kinetic friction coefcient well below 2%, and launch-
ing may therefore require pushing the deck to breakaway and braking it immediately after.

Bridge launching affects huge masses, and no remedies are available in case of uncontrolled sliding.
As a matter of fact, launching a deck along inclined surfaces requires prudence and experience.

2.11. Correction of launch stresses


The most characteristic aspect of the design of a launched bridge is the need to resist the stresses due
to the transient support congurations assumed by the deck during launch. Every cross-section of
the deck passes cyclically in midspan and above the piers, and is therefore subject to the maximum
positive moment, the maximum negative moment, and the maximum shear. Each cross-section has
to resist high self-weight transient stresses that are signicantly different from the service stresses,
and thermal stresses and the effects of geometry irregularities further complicate the situation.

The launch stresses can cause irreversible damage to the deck and accelerate deterioration over
time, and must therefore be analysed carefully and controlled with appropriate levels of launch
prestressing (AFGC, 1999). Steel girders are launched without prestressing, and stability and
structural detailing are the critical aspects of design. Bridge launching is indeed not as simple
as it appears, and it leads to a high-quality product only if bridge design and construction are
handled properly.

The state of stress in the deck evolves between the two limit conditions shown in Figure 2.39.
Considering two successive piers and a typical span of the continuous beam, the rst condition
is the nal position A, with the pier diaphragms of the deck located above the piers. The second
condition is position B, with the deck advanced by half a span and supported on cross-sections
that, once launching is complete, will be midspan sections.

In a PC bridge, in both positions, the cross-sections temporarily at midspan are rarely overloaded.
The shear is low, and bending is lower than the values reached under live loads. In addition, the
box girder cross-section is well suited to positive bending. The wide top slab provides a large com-
pressed area that draws the centroid upwards, and longitudinal prestressing is designed to control
the tensile axial stress at the bottom edge.

With regard to the support sections, in position A the exural launch stresses are only a fraction of
the service stresses, and the deck can be designed for the latter. Longitudinal prestressing is still

68

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


Bridge launching

Figure 2.39 Limit support conditions. (Reproduced with permission from ASCE)

needed to cover the tensile stress at the top edge. The webs are designed for the tangential stresses
in service conditions and checked for the launch stresses without the deviation forces of draped
tendons, which are installed only on launch completion.

In position B, cross-sections that on launch completion will be in midspan to resist positive


bending and minimal shear are subject to high shear and negative bending. They often need to
be adapted to these transient stress conditions, which makes them oversized with respect to the
service requirements. Every cross-section of the deck is a support section during launching, and
the need to resist the same transient stresses requires that moment of inertia and web thickness
be constant throughout the length of the bridge. This prevents lightening of the midspan sections
and further burdens the deck (Dezi et al., 1982; Rosignoli, 1996, 1997a, 1999c).

Uniform axial prestressing is provided during launching to alternately resist tensile stresses at the
opposite edges of the cross-section. The peak negative moment is about twice the peak positive
moment, the distance of the box girder centroid from the top bre is about one-third of the section
depth, the top bre modulus is about twice the bottom bre modulus, and a box girder is therefore
perfectly balanced for resisting the edge tensile stresses of launching with axial compression.
Internal or external draped tendons are added on launch completion to resist the service stresses.

Ribbed slabs with double-T section are also compatible with launching, provided that the webs are
designed for the longitudinal compressive stress due to the combined effects of negative bending
and axial launch prestressing in the support regions. If the cross-section is unable to contain all the
prestressing tendons, internal launch prestressing is combined with external draped tendons on
launch completion. Solid and voided slabs are typically used for 2025 m launch spans.

The behaviour of a prestressed composite box girder with steel corrugated-plate webs is more com-
plex during launching. Compared with a PC box girder, higher exural efciency and lighter
weight reduce the edge axial stresses of launching and require less launch prestressing. The launch

69

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


Bridge Launching

tendons are mostly located within the concrete slabs, aligning the prestressing force with the centre
of gravity of the cross-section is difcult with only internal tendons, and some external tendons are
often necessary. This results in more efcient nal prestressing schemes, as the external tendons
can be relieved and repositioned on launch completion. The external polygonal tendons of service
prestressing may be designed to balance the shear force due to self-weight and 50% of live loads
with tendon deviation forces, and the steel webs resist the shear uctuations due to the presence
or absence of live loads. During the launch, however, the steel corrugated-plate webs resist full
self-weight shear. This may result in oversized details, especially when the box girder is launched
without temporary piers.

Launching a non-prestressed composite bridge is simpler, as the steel girder is launched without
the concrete slab and its weight is therefore much lighter. The weight of a PC box girder, a
prestressed composite box girder and a non-prestressed composite box girder may be compared
for a typical 50 m launch span, a 13 m wide concrete slab and normal-weight concrete.

g A PC box girder with internal prestressing may have an average concrete thickness of about
0.60 m, and therefore weighs 195 kN/m.
g A prestressed composite box girder has an average concrete thickness of about 0.35 m and
a concrete weight of 114 kN/m. Adding 6 kN/m for two steel corrugated-plate webs, the
total weight becomes 120 kN/m, and the weight saving is 38%.
g The average thickness of the concrete slab of a composite box girder is about 0.28 m, which
results in a concrete weight of 91 kN/m. The steel U-girder weighs about 25 kN/m, the total
weight is 116 kN/m, and the weight saving is 41%. On 50 m spans, a composite box girder
is not much lighter than a prestressed composite box girder, and prestressing signicantly
decreases the weight of the steel webs.

Launching a composite deck complete with the concrete slab would offer signicant advantages in
terms of logistics, safety and impacts on the area beneath the bridge. However, the weight of the U-
girder is only 22% of the total weight of the cross-section. Launching the completed cross-section
would result in launch stresses ve times greater, and therefore the concrete slab is typically cast in-
place on completion of the launch of the U-girder.

The 25 kN/m weight of the U-girder is 13% of the 195 kN/m weight of the PC box girder. The
launch support reactions are much smaller, but the open U-girder is more exible, and large
exural rotations at the launch supports require rocking bearings. The centre of gravity of the
cross-section is often located below the middle of the girder depth, and this may result in oversized
top anges. Finally, the weight of the pier diaphragms causes peaks in the envelope of launch
bending.

Structures that present so many and different load conditions require careful pre-sizing, to avoid
excessive stresses in one of the many launch or service stages. Excessive prudence, however, would
result in an oversized structure. Most of the launch stresses depend linearly on the self-weight q of
the deck, and reducing this is of fundamental importance. Setting p as the distributed service load,
the cost of the deck depends on p + q, and the efciency rs of the structural design can be
expressed as

p
rs = (2.7)
p+q

70

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


Bridge launching

Weight reduction limits the inuence of the launch stresses and increases the efciency of the
design by creating a reserve available for service loads. The cost of structural materials and launch
equipment depends on the weight of the cross-section (i.e. on its area A). The moment of inertia I is
the main indicator of the exural capacity, and the efciency of the cross-section can be evaluated
in terms of stiffness-to-weight ratio (i.e. in terms of radius of gyration):

I
r= (2.8)
A

In a section the centre of gravity of which is a distance zu and zl , respectively, from the upper and
lower edge, the highest radius of gyration is achieved by concentrating masses at the edges:

r2max = zuzl (2.9)

and the cross-sectional efciency can be expressed as the exural efciency rf :

r2 I
rf = = (2.10)
r2max zu zl A

The closer rf is to the ideal value of 1, the better the exural efciency. In general, PC box girders
have rf 0.55, prestressed composite box girders with steel corrugated-plate webs can reach
rf 0.70 and non-prestressed composite sections may reach higher values.

In a continuous beam with constant self-weight and many spans of constant length L, far from the
ends the sections over the piers remain vertical, and the static system of each span is that of the
perfectly xed beam. During launching, however, the front region of the deck overhangs the entire
span prior to landing on the next pier. Negative bending in the deck section over the lead pier is six
times greater than the negative bending at the rear supports, and the shear is double. Immediately
after landing, the peak positive bending in the front span is 1.85 times greater than the midspan
bending in the rear spans. Designing a constant-depth deck for the transient launch stresses in the
front region would burden the entire bridge, and designing for the launch stresses in the rear region
would be inadequate for the front one.

The most logical and cost-effective solution is to design the deck for the service stresses and dimin-
ish the launch stresses in the front deck region, beyond that critical cantilever length Lcr = 0.41L
for which negative bending at the root of the cantilever is equal to the xity bending of the rear
region. There are three possible solutions, which may be adopted either individually or in
combination:

g limiting the launch stresses throughout the length of the deck by shortening the spans (i.e.
increasing the number of supports by inserting temporary piers between the nal piers)
g limiting the difference between the launch stresses in the front and rear deck regions by
supporting the front cantilever with a temporary adjustable cable-stayed system
g reducing the cantilever weight by applying a light steel extension to the front end of the
deck.

In the rst PC launched bridges, the launch spans were shortened by temporary piers distributed
throughout the length of the bridge. It was immediately evident that repositioning a temporary
pier span-by-span under the front cantilever involved signicant time and costs. Time and costs

71

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


Bridge Launching

were acceptable only on long spans, and long spans suggested launch stress reduction also in the
rear region of the deck. It was also observed that the use of multiple temporary piers in every span
would not produce the dramatic stress reduction predictable at rst sight, because of the different
exibility of the permanent and temporary piers, the construction tolerances and the exural stiff-
ness of the deck. Therefore, only one temporary pier is used in each span.

With the progressive increase in the cost of labour, temporary piers have been used less and less
frequently for the launch of PC bridges, being used only on long or varying spans. Their use is
infrequent also for the launch of steel girders, despite the possibility to cast the concrete slab
behind the abutment and launch the completed deck. Temporary piers are still the rst-choice
solution for control of overturning in the initial stages of launching, and for the launch of
cable-stayed bridges and decks suspended from arches.

Several steel girders have been launched by supporting the front cantilever with stay cables
anchored in the rear span and deviated by a mast supported on the girder. This scheme requires
continuous adjustment of the pull in the stays in relation to the position reached by the mast
during launching. Before landing at the next pier, the mast is behind the lead pier, and the stays
support the cantilever and reduce its deection. After landing, the pull is relieved so as not to
overload the front span with the load applied by the mast. The need for frequent pull adjustments,
the stress concentration at the base of the mast, the complexity of the operations and the risks in the
case of errors limit the use of front cable-stayed systems to steel girders on very long spans.

The third possibility is to limit the cantilever weight by means of a light extension, the launching
nose, which anticipates the landing at the next pier. The launch nose is such a safe, fast, efcient
and cost-effective solution that its adoption has become virtually standard in full-span launching
of PC decks. The reduction in the cantilever weight is advantageous also with the steel girders and
when temporary piers are used, and front stayed systems also take advantage of it. The launch
nose thus characterises most applications of incremental launching construction.

2.11.1 Launch noses


The lower the exural efciency (Equation 2.10) of the deck cross-section, the more important the
role of the launch nose. The steel girders of composite bridges are designed to resist the weight of
the concrete slab prior to the onset of composite action, and the exural capacity is so high during
launching that the launch nose may often be avoided or be used only for control of deections.
Heavier prestressed composite box girders typically require a launch nose, which may be obtained
by casting the bottom slab full-length in the casting yard, casting the top slab of the front span
after launch completion, and using the light front U-section as a launch nose. Heavy steel noses
are necessary for the launch of PC highway bridges, and the heaviest noses are used for the launch
of PC box girders for dual-track railway bridges.

Launch noses of all types have been used for 50 years steel trusses (Figure 2.40), braced built-up
I-girders, steel box girders, PC girders and prestressed composite girders in combination with
front realignment wedges or hydraulic systems for recovery from deection. Launch noses have
been custom designed or reused as found on the second-hand market.

The most common type of launch nose for single-cell PC box girders consists of two braced
steel I-girders connected to the front deck diaphragm. The girders are aligned with the bottom
webslab nodes of the deck in order to use the same launch bearings, and the soft of the nose

72

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


Bridge launching

Figure 2.40 Trussed launch nose with a 1500 mm deflection recovery system. (Photo: Torroja)

and deck is ush at the joint to enable continuous launching. When the deck launch surfaces are
inclined in the plane of the cross-section, the bottom anges of the nose have the same cross-fall
(Figure 2.41). The launch nose acts as a structural extension of the deck, and governs the exural
and shear stresses in the front region of the continuous beam.

Figure 2.41 Nose and deck with inclined launch surfaces. (Photo: URS)

73

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


Bridge Launching

Three dimensionless parameters govern the behaviour of the nosedeck elastic system:

g the length of the nose relative to the span to be overcome, Ln /L


g the weight of the nose relative to the self-weight of the deck, qn /q
g the exural stiffness of the nose relative to the exural stiffness of the deck, EnIn /EI.

The nosedeck interaction model discussed in the following section illustrates the inuence of
these three parameters, facilitates optimisation of the elastic system, and avoids trial-and-error
analysis when using more sophisticated means of calculation. More rened spreadsheet-based
design tools may be generated from the model by allowing for the varying weight and moment
of inertia of the nose, the weight of the front deck diaphragm, and the fact that the nose is typically
rectilinear also in bridges having a vertical curvature.

2.11.1.1 Nosedeck interaction


In a PC bridge, the launch nose may be pre-sized starting from a length Ln of about two-thirds of
the typical span (AFGC, 1999; Iglesias, 1992) and from a rst self-weight qn obtained using the
following equation:

qn = kL2n (2.11)

where (in kilo-newtons and metres) it is 0.012 , k , 0.02 for highway bridges and 0.018 , k , 0.03
for railway bridges, with progressively higher values in the case of wide and heavy PC box girders.
The following model is based on the following assumptions (Rosignoli, 1995, 1998c, 1999d).

g The nose and deck have constant stiffness and weight. As these actually vary, average
values may be used. The weight of the end deck diaphragm may be easily introduced
through the support rotations it causes. Thermal gradients, misaligned launch bearings and
geometry imperfections are disregarded in the analysis of nosedeck interaction as their
signs can reverse. The effects of irreversible creep deections may also be introduced if
necessary.
g The nose and deck have the same vertical curvature. When the deck prole is circular and
the launch nose is rectilinear, a varying vertical offset is applied to the front support in the
model. This effect is often negligible with the large vertical radii of highway and railway
bridges, but may be signicant in urban bridges having short vertical radii.
g The bridge has spans of constant length, and the number of spans behind support C is such
that the continuous beam behaves as if it were composed of an innite number of spans of
constant length L (Figure 2.42).
g Launch prestressing has a centroidal resultant throughout the length of the bridge and does
not cause hyperstatic exural effects.

The distance x of the nosedeck joint from pier B in Figure 2.42 denes the progression of launch.
The cantilever congurations before landing at pier A vary from the starting position, x = 0, to the
contact position, x = L Ln , and the dimensionless progress of launch

x
a= (2.12)
L

varies in the range 0 a , 1 Ln /L. At the start of launching, a = 0, the negative bending in the
deck section over pier B is that of the cantilever nose, which in a dimensionless form (the moments

74

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


Bridge launching

Figure 2.42 Nosedeck system in the first phase of launch

Launch direction

D C B A
L L L L

x Ln

due to cantilever masses will be marked with a point) is


 
MB 1 qn Ln 2
= (2.13)
qL2 2 q L

Once launching starts, bending increases (in absolute value):


 
MB a2 qn Ln 1 Ln
= a+ (2.14)
qL2 2 q L 2 L

This equation does not account for the weight of the end deck diaphragm (for the sake of this
discussion). The moment MC over pier C is

MC k5 + k4 1 k2 MB
= (2.15)
qL2 k3 + k1 qL2 k3 + k1 qL2

with the following coefcients

L
k1 = (2.16)
3EI
L
k2 = (2.17)
6EI
L
k3 =  (2.18)
2 3EI
qL3
k4 = (2.19)
24EI
qL3
k5 =  (2.20)
24 3EI

Finally, the support reaction at pier B is


 
RB MC (1 + a)2 qn Ln 1 Ln
= + + 1+a+ (2.21)
qL qL2 2 q L 2 L

75

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


Bridge Launching

On landing at pier A, the nose tip is jacked to allow the progress of launching on the launch bearings.
Deection recovery creates a positive moment that reduces MB , and the second launch stage starts,
in which the nose slides on pier A until the concrete deck arrives. This stage of launching is dened
by 1 Ln /L , a 1. The reaction RA at the front support is
   2
k5 + k4 1 k7 qn Ln
k2 k4 k8 + + a 1  
RA k + k1 qL2 2 q L a2 qn Ln 1 Ln
= 3 + + a + (2.22)
qL k22 2 q L 2 L
k1 + k6
k3 + k1

with the following additional coefcients


   
L EI
k6 = 1+ 1 (1 a)3
(2.23)
3EI En In

  
L 2 1 a EI 1 a2 a3
k7 = a + + (2.24)
EI 2 3 En In 6 2 3
  2 
qL3 q a a3 a3 a4
k8 = a(2 a) + n (1 a)2
2EI q 2 3 3 4
 
qL3 EI q  
+ 4a2 (1 a)3 + n 4a5 15a4 + 20a3 10a2 + 1 (2.25)
24EI En In q

and the value of MB in the second launch stage is

MB MB RA
= + (2.26)
qL2 qL2 qL

where MB is the cantilever moment (Equation 2.14) and the reaction RA intervenes only for
1 Ln /L a , 1. Equation 2.26 shows that, to be effective, a launch nose must reduce the
cantilever moment MB and at the same time ensure high values of RA . In other words, a launch
nose must satisfy two a priori contradictory requirements:

g it must be light, to reduce MB, and sufciently long that at landing at the pier, the heavy
cantilever portion of PC deck is as short as possible
g it must be stiff, so that deection recovery on landing produces an initially high reaction RA
that immediately increases, as launching proceeds, to balance the increase in MB caused by
the progressive lengthening of the PC deck section in the rst span.

For a long launch nose, Ln /L = 0.8, of relative weight qn /q = 0.1, Figure 2.43 describes the
progression of MB over the course of launching in relation to the exural stiffness. Prior to landing
at pier A, for a , 0.2, MB grows as per Equation 2.14. Upon reaching pier A, the positive moment
generated by deection recovery reduces MB, and as launch continues, MB tends to an end-of-
launch value MEOLB , which can be obtained by setting a = 1 in Equation 2.26:
 
MBEOL qn Ln 2
= 0.134 0.106 (2.27)
qL2 q L

76

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


Bridge launching

Figure 2.43 Progression of MB for Ln /L = 0.8 and qn /q = 0.1 in relation to EnIn /EI. (Reproduced with
permission from ASCE)

0.16

0.12
Bending at Pier B: M/ql 2

0.08

0.04 0.025 0.050


0.100 0.200
0.400 Fixity

0.00
0.00 0.25 0.50 0.75 1.00

In the cantilever conguration, a , 0.2, and in the nal stage of launch, a . 0.9, MB does not
depend on EnIn /EI. Therefore, the exural stiffness of the launch nose can be used only to prevent
the peak negative bending after landing and until stabilisation of the end-of-launch value (i.e. in
the launch phase 0.2 , a , 0.9) from exceeding the greater of the moments at landing and at the
end of launch. The optimum stiffness turns out to be around EnIn /EI = 0.2. Greater stiffness is
pointless, as it cannot prevent reaching the end-of-launch moment (Equation 2.27). Lower stiff-
ness increases the negative design moment in the front deck region. Finally, MB at landing is lower
than MEOL
B , landing is uselessly anticipated, and the launch nose may be shortened.

Figure 2.44 shows the progression of MB for a short nose (the same relative weight qn /q = 0.1, but
Ln /L = 0.5). Prior to landing, MB is affected favourably by the light cantilever weight of the nose,
but unfavourably by the long PC deck section overhanging beyond pier B. Upon landing and
deection recovery, MB is reduced effectively, and its value stabilises at the end of launch to
MEOL
B . The effects of the exural stiffness are also comparable, and in this case too the optimum
stiffness is around EnIn /EI = 0.2. However, keeping MB , MEOL B during the intermediate launch
stages loses its signicance, because this threshold is substantially exceeded in the cantilever con-
guration. With a short nose, MB at landing is higher than MEOL B , and the rst phase of launch
governs the design. Landing occurs too late (i.e. the nose is too short and should be extended).

Proceeding by trial and error, the optimum nose length is obtained as shown in Figure 2.45.
With EnIn /EI = 0.2, MB at landing is equal to MEOL B , and the peak negative bending,
MB,max = 0.101qL2, is 22% higher than the xity bending in the rear deck region. With a exible
nose with EnIn /EI = 0.025, in spite of the correct nose length, negative bending is out of control in
the intermediate launch stages, and the peak negative bending, MB,max = 0.148qL2, is 78%
higher than in the rear deck region.

Figure 2.45 shows the optimum diagram of MB for qn /q = 0.1. Comparisons of diagrams of MB
optimised for different values of qn /q show that MB depends signicantly on the nose weight only

77

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


Bridge Launching

Figure 2.44 Progression of MB for Ln /L = 0.5 and qn /q = 0.1 in relation to EnIn /EI. (Reproduced with
permission from ASCE)

0.16

0.12
Bending at Pier B: M/ql 2

0.08

0.04 0.025 0.050


0.100 0.200
0.400 Fixity

0.00
0.00 0.25 0.50 0.75 1.00

during the rst phase of the launch. Because of the different relationship between MB and qn /q in
the two phases of launch, once the system has been optimised for a given relative weight, an
increase in qn /q modies the diagram of MB toward the type shown in Figure 2.44 (short nose),
and a reduction modies the diagram toward the type shown in Figure 2.43 (long nose).

To force MB at landing toward MEOL B it is necessary to shorten the cantilever section of the PC
deck (i.e. to lengthen the nose) as qn /q increases. The diagrams of MB that restore this condition
are compared in Figure 2.46. Starting from qn /q = 0.1, as the relative weight increases, equalisa-
tion of MB at landing and MEOL
B is obtained with longer noses and for progressively lower values

Figure 2.45 With qn /q = 0.1, negative bending at landing is equal to the end-of-launch moment for
Ln /L = 0.65. (Reproduced with permission from ASCE)

0.16

0.12
Bending at Pier B: M/ql 2

0.08

0.04 0.025 0.050


0.100 0.200
0.400 Fixity

0.00
0.00 0.25 0.50 0.75 1.00

78

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


Bridge launching

Figure 2.46 Negative bending is equalised with the nose length as qn /q varies. (Reproduced with
permission from ASCE)

0.12
Bending at Pier B: M/ql 2

0.08

0.04
0.050 0.100
0.150 0.175
Fixity

0.00
0.00 0.25 0.50 0.75 1.00

of MB , which tend to the xed beam moment of the rear deck region. Proceeding by trial and
error, Figure 2.47 denes the optimum values of Ln /L in relation to qn /q for the optimum exural
stiffness. More precise evaluations can be made by considering that variations in the nose weight
affect the exural stiffness, and vice versa (Granata et al., 2013).

The diagram of negative bending at pier C shows that the maximum of MC in the rst phase of
launch is always reached for a = 0. Independent of the nose length, deection recovery at pier
A forces MC toward a single curve that stabilises for a = 0.6 on a value slightly lower than the
xity moment. From this point on, negative bending in the rear deck region is no longer inuenced
by the front nosedeck system. With regard to negative bending, therefore, the use of a launch

Figure 2.47 Optimal relative length Ln /L for negative bending as qn /q varies. (Reproduced with
permission from ASCE)

1.000
Relative nose length

0.750

0.500
0.020 0.060 0.100 0.140 0.180
Relative nose weight

79

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


Bridge Launching

nose of sufcient stiffness and adequate length for the relative weight limits the deck zone
inuenced by the nosedeck interaction to a length of about 1.5L, increasing to 2L in the case
of very deformable noses.

The maximum positive moment in the front span is


      
MAB,max RA qn Ln 1 RA 2 1 qn Ln 2 qn
= 1 a + + 1 (2.28)
qL2 qL q L 2 qL 2 q L q

In contrast to what happens to MB , the exural stiffness of the launch nose affects only the
progression of MAB,max toward the maximum value, which is, in any event, reached for a 0.9.
It then starts to decrease because of the boundary restraint provided by MA. The progression
of MAB,max with different nose lengths is substantially similar. The peak positive moment depends
essentially on the continuous-beam behaviour of the PC deck, and can be modied only by means
of MA. Increasing the length and the weight of the nose can, therefore, diminish the negative and
positive design bending in the front deck region. Finally, the consideration that the end span of a
continuous beam is always subject to higher positive bending is misleading. The end spans are
typically shorter than the interior spans to compensate for these increases, while during launching
the front deck region has to overcome all the longer interior spans.

For the BC span one obtains


 
MBC,max 1 MC MB 1 2 MB
= + + 2 (2.29)
qL2 2 qL2 qL2 2 qL

The diagram of MBC,max conrms that the rear deck region unaffected by the nosedeck inter-
action begins at a distance of about 1.5L from the nosedeck joint. In conclusion, the nosedeck
system may be designed as follows:

1 Start from a nose weight obtained from Equation 2.11.


2 Consider the localised weight of the end deck diaphragm.
3 Rene the value of Ln /L until MB at landing is equal to MEOL
B .
4 Determine the minimum value for EnIn /EI to keep MB below this threshold in the
intermediate launch stages with 1 Ln /L , a 1.
5 Design the launch nose for the EnIn /EI ratio, and adjust the qn /q ratio.
6 Iterate to convergence.

The iterations are simple to perform with a spreadsheet. An optimisation algorithm has been
proposed for the nosedeck interaction. It is based on the method of feasible directions, and
handles sets of xed parameters, design variables and objective functions (Fontan et al., 2011).
The algorithm allows the choice of multiple optimisation parameters (e.g. MB at landing equal
to MEOL
B and MC,max , or equal peak tensile stress at the top and bottom bres under negative and
positive bending for optimal design of launch prestressing), considers the relationships between
nose stiffness and weight, and prevents physically impossible solutions (e.g. a nose that is too light
for the required exural stiffness).

2.11.1.2 Nose design


The optimum length and stiffness of a launch nose depend on many factors. Depreciating the
investment on several projects suggests designing the launch nose to be as versatile as possible.

80

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


Bridge launching

PC bridges may be launched on spans up to 6570 m, and the nose may be designed for the longest
spans and modied for the specic application by shortening the rear segment.

Most launch noses for PC bridges include two built-up steel I-girders composed of transportable
segments connected by lateral bracing and cross-frames at the nodes of lateral bracing. Box noses
have also been used, but are typically more expensive. PC noses that are to be demolished on
launch completion have been used in the case of difcult transport. Composite noses comprising
steel webs and PC anges have been used for the launch of heavy decks. Truss noses are light and
efcient, but the demand for hand welding and the weight of bottom chords designed for the
migration of support reactions make their use infrequent for the launch of PC bridges. Truss noses
are often the best solution for the launch of lighter steel girders.

When custom designed for the bridge, the rear depth of the nose is equal to the deck depth, and is
maintained for a few metres from the joint section. Controlling negative bending in the deck
requires full exural stiffness of the nose only in the last launch phases, and a lighter front section
saves costs and cantilever mass.

Cross-frames and lateral bracing connect the two I-girders from common nodes. The cross-frames
brace the anges and control the load difference between the main girders; when the nose is
twisted, they also oppose distortion. The cross-frames are designed as crosses, K-frames, or
cross-beams that connect vertical web stiffeners in the plane of lateral bracing (Figure 2.48).
The cross-beams may be stiffened with diagonals to further stabilise the top ange. The walkway
to the front working platform may be supported on the cross-beams.

Figure 2.48 Cross-beams framed to web stiffeners

81

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


Bridge Launching

The role of lateral bracing is to resist lateral wind and the loads applied by the launch guides.
Lateral bracing is designed with crosses between the bottom anges or with a midplane truss where
the cross-beams are the verticals. One plane of lateral braces is sometimes used in rectilinear
bridges to reduce the cost of adapting to decks of different widths; curved decks typically require
bracing of both anges for enhanced torsional stiffness and strength. In narrow noses, the cross-
frames are spaced so as to achieve square panels for the lateral bracing; rectangular panels are
more common in wider noses.

The design of a launch nose starts from the envelopes of bending and shear determined in the
launch stress analysis. When the deck has a circular prole, additional bending and shear derive
from the displacement of a rectilinear nose along the cylindrical launch alignment. When the deck
is curved in plan, the nose is also subject to torsion and uneven shear forces in the webs. Design
envelopes are calculated for the design-governing I-girder, the web and anges are designed for
longitudinal bending and shear, and local stiffeners are applied where necessary to resist local
effects.

The design stresses are kept reasonably low to accommodate local effects and load differences
between the two I-girders due to fabrication and assembly tolerances, atness defects in the
bottom anges, geometry imperfections in the deck and misaligned launch bearings. The load
differences may be signicant in launch noses having closely spaced girders, and the cross-frames
are typically too exible to fully control them. Load imbalance is minor at deection recovery
because of the nose exibility, grows during the course of launching, and is maximum when the
deck reaches the pier because of the high torsional stiffness of a PC box girder. Each I-girder
of the nose is, therefore, designed for 75100% of the whole support reaction. This also applies
to the prestressing systems that connect the nose to the deck. The PC box girder is also checked
for the resulting torsion (AFGC, 1999).

The breadth of the bottom ange is chosen so that the same Neoon pads can be used for the
nose and the deck; 300 mm is a popular width. Transverse bending governs the thickness tb of the
bottom ange. The peak transverse axial stress generated by the support reaction RA is calculated
assuming that the entire support reaction is applied to one girder only:

3RA bnf
st = (2.30)
4t2b SLnf

where bnf is the transverse breadth of a Neoon pad, Lnf is its length and SLnf is the total length of
the Neoon pads on a launch bearing. Once this rst-trial bottom ange area has been determined,
the area of both anges is calculated by trial and error, based on the expected web thickness, the
resulting location of the cross-section centroid, the serviceability limit state (SLS) stress limits speci-
ed by the design standards, and the ultimate limit state (ULS) cross-sectional capacity as identied
by the design standards for bending and shear in relation to the cross-sectional compactness. The use
of hybrid sections including different grades of steel is infrequent. The nal dimensions of the ange
plates are checked against the breadth-to-thickness ratio specied by the design standards for
control of instability.

The web thickness tw is determined from the envelope of the shear force. The tangential stress is
kept low to account for the uncertain load distribution between the girders and to distribute the
vertical web stiffeners according to the bracing geometry. The web thickness is rarely less than

82

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


Bridge launching

Figure 2.49 Vertical flange stiffeners. (Reproduced with permission from ASCE)

12 mm to facilitate workshop handling; most often it is 20 mm or thicker. Local buckling of the


web panels is checked for combined banding and shear.

By assuming 458 dispersal of the support reaction within the bottom ange and the entire support
reaction applied to one girder only, the vertical axial stress at the bottom of the web panel over the
launch bearing is

RA
sz =   (2.31)
tw SLnf + 2tb

If the transverse axial stress in the bottom ange or the vertical axial stress in the web panel is
excessive, closely spaced ange stiffeners can be used to create additional support lines in the
ange and to improve the vertical load dispersal within the webs (Figure 2.49). Flange stiffeners
are typically necessary only in the rear section of the nose. The cost of hand welding discourages
the use of a great number of ange stiffeners.

Triangular corner stiffeners welded to the ange and web on both sides do not require hand weld-
ing but complicate the use of longitudinal stressed bars in the eld splices due to jacking clearances
(Figure 2.50). Corner stiffeners increase the torsional stiffness of the webange node, and provide
local beam action and direct longitudinal load transfer between the vertical web stiffeners, which
enhances the dispersal of support reactions (Mato and Santos, 2007).

When the nosedeck joint is close to the pier, the rear region of the nose is severely stressed by the
concomitance of multiple load-transfer mechanisms:

g The support reaction and the nose portion overhanging beyond the pier generate
longitudinal bending and shear.

83

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


Bridge Launching

Figure 2.50 Triangular corner stiffener. (Reproduced with permission from ASCE)

g Dispersal of the support reaction within the webs generates vertical axial stress.
g Strut-and-tie mechanisms arise from the application point of the support reaction to the
shear keys welded to the end plate of the nose. When a large shear key is applied at the
bottom of the end plate, the transfer of shear forces is more localised (Figure 2.51).
g The anchor systems of the prestressing bars that clamp the nose to the deck generate highly
localised stresses, and create stiff spots that collect axial and tangential stresses from general
bending and shear.

The principal stresses may be calculated assuming a biaxial state of stress; the use of three-
dimensional nite-element analysis is necessary in the most delicate cases. The complexity of the
boundary conditions suggests the need to model the entire joint region: the rear portion of
the nose, the end diaphragm of the deck, the front portion of the box girder, the anchor bars and
the launch prestressing tendons. Symmetry about the deck centreline may help to simplify the
model and shorten the runtime.

When ange stiffeners or triangular corner stiffeners prevent the operation of the stressing jack for
the anchor bars, the bars are tensioned from the rear anchorages within the box cell. Four anchor
bars are often used on each layer to enhance the exural efciency of joint prestressing; their anchor
boxes are welded to webs and anges for direct load transfer (Figure 2.52). Three-dimensional nite-
element analysis facilitates nose detailing and the design of transverse prestressing in the end
diaphragm of the deck, especially when the deck has inclined webs.

Longitudinal stressed bars are frequently used for the eld splices in both anges (Rosignoli,
2013). At the bottom ange, the use of lap plates for slip-critical bolted connections requires lifting

84

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


Bridge launching

Figure 2.51 End plate and bottom shear key

Figure 2.52 Anchor boxes for the joint prestressing bars

85

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


Bridge Launching

of the deck to avoid interference with the launch bearings. When a steel girder is launched on
rollers located under the webs, two narrow lap plates may be used at the bottom ange to allow
the launch rollers to pass through. The splice geometry is the same for the launch nose and the
main girder in order to be able to use the same rollers for launching.

The use of stressed bars optimises the contribution of the different elements of the connection:
controlled-tightening bolts transfer shear by friction, and longitudinal stressed bars anchored to
the anges transfer bending by means of a post-tensioning system designed for load reversal.
The bars are lodged into boxes welded to web and ange, the welds transfer longitudinal axial
force from the ange to the bar anchorage, and the absence of lateral contact between the bar and
the containment box avoids shear transfer.

After tightening the shear bolts, the bars are inserted into the anchor boxes for symmetrical ten-
sioning. The absence of bolts and lap plates at the bottom ange leaves the nose intrados ush for
continuous launch operations. The web may be spliced using a lap-plate connection or a frontal
splice with an end plate. The end plate acts as a vertical web stiffener, longitudinal web stiffeners
are welded to the end plate and spliced without bolts, and the shear force applied to the bolts is
independent of their distance from the centre of rotation of the splice. The stiffness of this splice
arrangement is similar to the stiffness of a welded connection.

Not many standards govern the design of frontal splices that combine large-diameter stressed
bars and shear bolts. The bars are stressed by torque or by using hollow plunger jacks; the use
of aluminium jacks is recommended for easier handling, and double-acting operation permits
fast retraction. For torque tensioning, FEM 1.001 (FEM, 1987) denes the tangential stress tb
generated by torque as
 
2db,min sb pt
tb = + 1.155mt (2.32)
db pdb,min

where db is the diameter of the bar, db,min is the minimum diameter, sb is the axial stress generated
by tightening, pt is the pitch of the thread, and mt is the friction coefcient of the thread. The ideal
stress at bar tightening is assessed using the following equation:

sid = s2b + 3t2b 0.8fy (2.33)

The initial stressing must be accurate, the bars can be re-tensioned when necessary, and a load
tolerance of +10% can be easily maintained over time. A resistance factor f = 0.9 may therefore
be used for the ULS design of the splice. When the bars are tensioned using a hollow plunger jack
and no torque is applied, tb = 0 is used in Equation 2.33.

The dimensions of the stressing jacks and anchor nuts govern the geometry of the connection. The
anchorages of multiple bars may be staggered longitudinally to reduce the concentration of stress
and to facilitate transfer of the axial force from the ange plate to the anchor boxes. The pre-
stressing force in the bars is designed to avoid joint decompression at the SLS. FEM 1.001
(FEM, 1987) species not exceeding the yield strength of steel with load factors g1 and g2 , and
avoidance of joint decompression with load factors g2 and g3 and a resistance factor f = 0.9.
The rst load factor depends on the conditions of the contact surfaces, and for attened end plates
it can be taken as g1 = 1. The second load factor is 1.1 , g2 1.5 in relation to the probability of

86

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


Bridge launching

the different load combinations. The third factor is 1 g3 1.3 and determines the margin from
joint decompression.

The load factors g2 and g3 are applied to the least favourable condition in relation to the time-
dependent relaxation of the tension in the steel, and conservative load distribution models are
applied between the adjacent I-girders. For the yield checks, an equivalent cross-sectional area
of the assembly is determined using the following equation:

 2
p L
Aa,eq = dcs + a dh2 (2.34)
4 10

where dcs is the diameter of the contact surface of the bar anchor nut, La is the total nut-to-nut
length of the assembly and dh is the diameter of the hole lodging the bar. The equivalent cross-
sectional area Aa,eq is used to compute the shortening DLa of the assembly under the tightening
force. Bar elongation DLb at tensioning is calculated taking into account the threaded end portions
of the bar (net area) and the central non-threaded portion:

DLa
db = (2.35)
DLa + DLb

The net area Ab of the bar is determined for db,min and the axial stress limit is specied by the
design standards for the SLS:

sb,max = fy2 3t2b (2.36)

Setting Fb as the longitudinal tensile force that the bending moment in the splice applies to the bar,
yielding is checked using the following equation:

Fb sb,max sb
(2.37)
Ab g1 g2 db

and joint opening is checked using

Fb fsb
(2.38)
Ab g2 g3 (1 db )

Stressed bars are also used for splicing the chords of trussed noses. Transfer of axial tension
governs the design of chord splices. Bar anchor pipes are welded symmetrically to the anges
and the web to streamline the stress ow from the tension chord to the bars and to increase the
robustness and redundancy of the splice. The use of several bars with staggered anchor nuts
simplies welding and often permits the use of commercially available prestressing bars. Local
bending and shear due to the migration of the support reactions along the bottom chord are
checked using the criteria discussed above. The end plates may be equipped with spindles that
act as shear keys and help keep the end plates aligned during assembly to avoid involuntary
eccentricities.

Frontal splices with stressed bars and shear bolts are structurally efcient and quick to assemble.
Because of the high capacity of the bars, however, these splices must be designed and tightened
with care.

87

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


Bridge Launching

g The distribution of bending and shear between chords and braced I-girders is often uneven,
and the splices are designed to allow for the possibility of accidental overloading.
g The use of two bars at every ange is a non-redundant detail. Failure of one bar may lead
to collapse due to local bending generated by load eccentricity and the absence of
alternative load paths and stress redistribution. Bars, anchor boxes and welds are
oversized to allow local damage at bar failure but to avoid collapse. Wide ange plates
can accommodate two anchor boxes on either side for enhanced redundancy. Narrow
anges accommodate only one box on either side, but a second bar may be lodged in a
round pipe welded to the edge of the ange, or a second anchor box may be welded to
the web.
g Stiff anchor boxes are used to protect the bars from impacts. Thick anchor plates and
short-pitch threads diminish the prestress losses due to anchor set. The central portion of
the bars is not threaded in order to save machining costs and to increase the shear capacity
of the bar in case of accidental shear transfer due to splice slippage.
g The use of calibrated bolts for bearing shear transfer is not recommended, as splice slippage
at load reversal may apply shear forces to the bars. Symmetrical hole patterns, the use of
large-diameter bolts, the conservative friction coefcients specied by the design standards
and the additional compression generated by the stressed bars typically lead to eld splices
that have a low demand-to-capacity ratio for shear and large margins from slippage. The
bolts are tightened concentrically from the centre of the splice, and the bars are inserted
into the anchor boxes on completion of bolt tightening.
g Accurate attening and sandblasting of the contact surfaces are recommended along with
the use of direct tension indicators on the bolts. Sealing the perimeter of the splice with
paint at the end of tightening is also recommended, as corrosion of contact surfaces can
diminish friction.

Lateral bracing and cross-frames are designed to align the nose girders with the bottom webslab
nodes of the deck to use the same launch bearings. Wider launch bearings permit the reuse of
narrow launch noses, but the lateral eccentricity of the load applied to the deck often requires
transverse post-tensioning in the end deck diaphragm. When buying a second-hand nose, there-
fore, some extent of retrotting should be budgeted.

In order to use only one set of lateral guides for launching, the outer edge of the nose bottom
ange may be aligned with the edge of the PC deck. Guiding the nose is only aimed at aligning
the support reactions with the nose webs. General deck alignment is corrected at the lead pier
reached by the PC deck. Attempting this operation on the nose will likely result in damage to nose
bracing.

2.11.1.3 Connection to the deck


The connection between the launch nose and the deck ensures transfer of bending and shear under
full structural continuity. The connection must be simple to assemble and to control during
launching. On launch completion, the connection must also be easy to dismantle for removal of
the launch nose.

In PC bridges, the most effective solution for the transfer of bending and shear is joint prestressing
with high-strength bars. The use of strand is less frequent due to difcult release and the larger
anchor plates. The bars are embedded in the deck for 1.52 times the deck depth to obtain an
anchor region as long as the deck depth both for the anchor bars and for the deck launch tendons

88

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


Bridge launching

Figure 2.53 Side view of the nose connection systems. (Reproduced with permission from ASCE)

(1.5 2)H

H
Shear keys

Temporary concrete segment

(Figure 2.53). As positive and negative bending in the nose are primarily resisted by the anges of
the I-girders, the anchor bars are divided into two bundles, and the resultant force of the two
bundles is kept as close as possible to the anges for direct transfer of axial load.

The peak positive bending due to travelling support reactions is typically higher than the negative
bending due to the cantilever nose. Most of the anchor bars are therefore located at the bottom
ange level. Some launch noses have been designed with a U-section in the joint region so that
multiple anchor bars can be used in the bottom slab to facilitate overlapping with the deck launch
tendons. In most cases, the nose is made with two braced I-girders, and the distribution of the
anchor bars is governed by technological restraints such as the nose geometry and the dimensions
of the bar stressing jack.

The rear anchorages of the anchor bars can be embedded in the deck or can be made accessible.
When the depth of the deck allows comfortable work within the box cell, most of the anchor bars
are tensioned from the rear anchorages. This facilitates inspection and if necessary re-tensioning,
the bar geometry is less restrained, and the front bar anchorages may be closer to the nose webs
because no jacking clearance is needed on the nose side. This results in compact joints and no risk
of workers falling when checking and re-tensioning the bars (Figure 2.54).

The most practical solution to transfer longitudinal prestressing from the anchor bars to the
launch tendons consists of modifying the layout of the launch tendons so as to introduce them
deeply among the bars. The non-contact splices between tendons and bars generate strut-and-tie
mechanisms that include diagonal compression struts in the concrete and transverse tension ties in
the rebar. Some launch tendons must be kept within the slabs to avoid excessive congestion at the
webslab nodes, and this portion of launch prestressing is transferred by means of longitudinal

89

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


Bridge Launching

Figure 2.54 Bar tensioning from the rear anchorages leads to compact end plates

interface shear at the webslab nodes. This often requires transverse prestressing in both slabs for
control of cracking.

Splicing of joint prestressing may be complex also in the top slab, especially when inclined deck
webs move the upper webslab node away from the top ange of the nose. This can require mild
transverse prestressing in the top slab to resist the principal tensile stress of interface shear transfer.
The top ange of the nose may also be anchored to an RC block match-cast onto the top slab. This
solution is often necessary in the presence of many launch tendons in the top slab, when the deck
webs are inclined, or when a second-hand nose is deeper than the deck (Figure 2.55).

On the nose side, the bars are symmetrical about the webs and are anchored to special vertical
stiffeners located 0.30.5 m ahead of the end plate. Horizontal stiffeners support the vertical
stiffeners from the end plate and transfer longitudinal compression (Figure 2.54). The load
transfer mechanism from the nose anges to the anchor bars includes welding of the web to the
ange, vertical anchor plates and horizontal stiffeners. The vertical weld between the web and the
end plate is mostly unaffected, and this provides robustness and redundancy to the connection.

The launch procedures include periodic checks of the load in the anchor bars and re-tensioning
them if necessary; the anchor bars are therefore left non-grouted and protected from stress
corrosion with passivating wax. The use of load cells on a few bar anchorages is recommended
in long bridges for earlier detection of excessive time-dependent losses of prestress.

Starting from the peak positive and negative bending in the nosedeck joint and from the distance
between the centroid of the deck cross-section and the resultant force of the upper and lower

90

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


Bridge launching

Figure 2.55 Front view of the anchor systems in Figure 2.53. (Reproduced with permission from ASCE)

Launching nose outline

Shear keys

bundles of bars (eu and el , respectively), the calculation of the lower prestressing force Fl and of the
upper prestressing force Fu is immediate. The anchor bars are oversized to resist load differences
between the nose girders without joint decompression, bar overstress and prying action.
Allowance is also made for re-entry of anchorages, friction losses and time-dependent losses of
prestress, as launching of a long bridge may take several months.

Negative bending due to the cantilever nose is given by Equation 2.13. Net positive bending after
landing varies over the course of launching:
 
Mj RA 1 qn Ln 2
= ( 1 a) (2.39)
qL2 qL 2 q L

and peaks at a 0.70.8 in relation to the relative stiffness of the nose (Figure 2.56). Equations
2.13 and 2.39 refer to the total moment in the two girders. To account for the irregular load

Figure 2.56 Bending in the nosedeck joint for Ln /L = 0.65 and qn /q = 0.1 in relation to EnIn /EI

0.06
0.025 0.050
0.100 0.200
0.400
Joint bending: M/pl 2

0.03

0.00

0.03
0.00 0.25 0.50 0.75 1.00

91

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


Bridge Launching

distribution, the nosedeck joint is designed so that one girder can resist most of the positive
bending. The distribution factor k varies between 0.5 (uniform distribution) and 1 (only one girder
resists positive bending). The end plate of one nose girder, of net area Aj (bar holes detracted) and
moment of inertia Ij , is designed under the assumption of linear distribution of the contact stress
along the joint surface. The edge stresses for maximum positive bending


Fu + Fl kMj + Fu eu Fl el
su =
+ zu
Aj Ij
(2.40)

Fu + Fl kMj + Fu eu Fl el

l
s = z l
Aj Ij

are compared with the SLS compressive stress limit specied by the design standards and the desired
least joint compression. The longitudinal compression due to the frictional resistance of the launch
bearings is usually disregarded. The same checks are performed for the negative bending generated
by the cantilever nose and the downward eccentricity of joint prestressing; for this check, bending
and axial force are spread uniformly between the two girders. Finally, the same checks are
performed using factored loads and resistances for the load combinations specied for the ULS.

Vertical shear peaks when the PC deck is reaching the pier. Most of the support reaction is
assumed to be applied to one girder only, and the nose web and web reinforcement in the deck are
designed accordingly. Large shear teeth at the bottom of the end plate or multiple shear keys
distributed along the plate are used for shear transfer; the shear capacity of the anchor bars is
considered only at the ULS. In spite of the presence of shear keys, joint prestressing is checked for
shear transfer by friction with a friction coefcient Cf = 0.4 for the steelconcrete interface and a
resistance factor f = 0.5. Vertical shear may also be transferred with steel brackets supported on
the top slab and anchored with vertical bars (Figure 2.57).

When the launch nose is deeper than the deck, the top anges are anchored by means of RC blocks
match-cast over the top slab. Reinforcement is rarely used for the connection to avoid demolishing
the block, cutting the rebar and patching the deck surface on launch completion. The RC blocks
are anchored to the deck by friction. The longitudinal force that the top ange applies to the block
acts as a shear force in the horizontal joint. The vertical force to be applied by means of pre-
stressing bars is determined with a friction coefcient Cf = 0.6 between two concrete surfaces and
a resistance factor f = 0.5. The equivalent friction coefcient can be increased with shear keys
obtained by pressing corrugated sheet metal into the fresh concrete cover of the top slab. The
friction connection of steel anchor blocks is designed with Cf = 0.4, and the bottom surface of the
block is equipped with shear keys (Figure 2.57).

Similar considerations govern the design of the connection between the launch nose and a steel
girder. The launch nose for a U-girder with inclined webs has a rear diaphragm that transitions
the nose girders into the inclined webs. The joint may be designed as a slip-critical connection with
lap plates. Bottom anges and chords are typically spliced with longitudinal stressed bars to keep
the bottom surface ush and avoid interference with the launch rollers. When the eld splices in
the steel girder are welded, a welded splice is often the most efcient solution also for the nose
deck connection (Figure 2.58).

Because of its critical role in the control of the launch stresses, the launch nose is load tested for
strength and stiffness prior to the rst launch. After assembling the nose and tensioning the anchor

92

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


Bridge launching

Figure 2.57 View of the anchor blocks in Figure 2.49 from inside the casting cell

Figure 2.58 Joint diaphragm between the truss nose and the U-girder

93

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


Bridge Launching

bars, the nose tip is jacked to create a positive moment in the nosedeck joint 2030% higher than
the design moment, and the tip deections are surveyed during loading and unloading for linearity
checks.

2.11.1.4 Recovery of elastic deflection


When approaching a new pier, the nose tip is generally distant from the theoretical launch align-
ment because of the exibility of the nosedeck elastic system. When the spans have similar length
and the piers have similar axial stiffness, the nose deection is directed downward at landing, and
it may be 0.10.2 m in a PC bridge and 1.0 m or more in a steel girder. Spans of different length
and exible temporary piers alternated with stiffer permanent piers may also cause upward deec-
tions at landing, which requires jacking or shimming of the launch bearings on the new pier to
achieve the required support reactions.

In most cases, the nose deection is directed downward. The need to recover this deection to
continue the launch is sometimes solved by inclining the bottom anges in the front region of the
nose, so that their advance on the pier forces progressive realignment. This alignment method is
used rarely and only for light steel girders, as the longitudinal force generated by the front wedge
can overload the piers. The launch rollers are anchored to the pier cap to avoid displacement or
overturning, and are articulated to cope with large rotations of the support surfaces.

For a PC deck, the simplest solution consists in using lifting jacks at every pier cap. The jacks are
positioned behind the launch bearings and support small sliding plates. When the launch nose
arrives over the jacks, Neoon pads are inserted between the jacks and the nose, the jacks are
extended to realign the nose, and launch may continue over the launch bearings. This solution
requires transferring the lifting jacks from pier to pier, and the short stroke of at jacks may
require several lifting cycles for deection recovery, which delays restarting of the launch.

When the bridge is long and multiple realignment operations are therefore necessary, vertical long-
stroke double-acting hydraulic cylinders are applied to the tip of the nose. The cylinders may be
equipped with a bottom plate that takes support onto Neoon pads over the launch bearings, and
the plungers are extended to lift the nose tip to the launch elevation (Figure 2.40). The cylinders
may also be used to rotate a rocking arm pivoted to the nose tip and also supported on the Neoon
pads (Figure 2.35). In either case, the realignment system advances with the nose and is immedi-
ately operational at the next landing.

2.11.2 Front cable-stayed system


An alternative solution to reducing the stresses from a long front cantilever consists of supporting
the tip of the cantilever with a temporary cable-stayed system comprising a mast supported on the
deck and strand cables fan-anchored symmetrically to obtain an adjustable external prestressing
system of high eccentricity (Figure 2.59). This solution is used more frequently for the steel girders
than for PC decks; its rst application to a PC deck dates back to 1969, on the Boivre Bridge in
France (Figure 2.60) (AFGC, 1999).

In the initial phases of launching, when the two leading spans of the deck are still in the casting
yard, a short launch nose and a steel mast are applied to the deck. The mast is positioned so that
when the nose tip reaches the landing pier the mast is a few metres behind the lead pier. The mast
takes support over the deck webs and is equipped with long-stroke support cylinders that control
the load applied to the deck and the pull in the stay cables.

94

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


Bridge launching

Figure 2.59 Temporary front cable-stayed system. (Photo: LaViolette)

For the system to be effective in all launch phases, it is necessary to modify continuously the pull in
the stays because of the load that the mast applies to the deck. When the nose tip approaches the
landing pier, the load is at its highest and the system is fully effective (Figure 2.60). When the mast
passes over the lead pier and moves toward midspan, the load applied to the deck will increase the
positive moment in the front span, and must be reduced by loosening the stay cables.

Figure 2.60 Position of maximum pull in the stays of the Boivre Bridge. (Reproduced with permission
from ASCE)

95

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


Bridge Launching

In PC bridges, several factors limit the use of front cable-stayed systems to the longest spans, and
always in combination with a launch nose that reduces the cantilever weight and the load in the
mast and stay cables.

g Before cantilevering out from the abutment, the deck must be long enough to anchor the
stay cables, and the casting yard must therefore be as long as two spans. Using a temporary
pier in the end span of the bridge may shorten the casting yard but increases the cost of the
facility and the labour demand during launching.
g The landing pier must be reached during each launch to avoid long launch stops with
stressed stays. This suggests launching deck sections as long as an entire span. Control of
creep deections requires ending each launch with a front cantilever to reduce positive
bending in the lead span, which adds complexity to the casting process. Using a long nose
and half-span launches simplies the design of the casting yard and diminishes the
accumulation of creep deections, but increases the cost of the nose (Figure 2.61).
g Structural analysis is quite complex. The pull in the stay cables depends on the deections
of deck and permanent and temporary piers, on bearing misalignment and geometry
irregularities, and on thermal deformations in the deck and the stay cables. Intrinsic
structural delicateness and the risks in case of error require higher safety factors in the
design of the system, and the nonlinear behaviour of the stay cables reected in the Ernst
modulus depends on the stress level, and may be pronounced because of the low pull
resulting from high safety factors.
g A long front deck region requires additional launch prestressing, and the stress concentration
in the deck region under the mast may cause cracking and requires additional reinforcement.

Figure 2.61 Launch of the Charix Viaduct on 64 m spans

96

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


Bridge launching

g The need for continuous control and adjustment of the pull in the stay cables slows
launching down and increases the demand for skilled labour.

For these reasons, the combined use of a short launch nose of length 0.3 , Ln /L , 0.5 and of
a steel mast of height 0.5 , hm /(L Ln ) , 0.8 is limited, in PC bridges, to wide and heavy box
girders with 6070 m spans, to solid, voided or ribbed slabs unable to resist the cantilever weight
of a long launch nose, and to curved bridges to diminish the lateral displacements of a rectilinear
launch nose on the launch bearings. Other than in these circumstances, long PC spans are typically
launched with one temporary pier per span.

The use of front cable-stayed systems may also be advantageous when a long central span is
obtained by launching two decks symmetrically from the opposite abutments for midspan closure.
The cable-stayed support of the two cantilevers permits reaching a central span much longer than
100 m. If the central span is designed with extradosal prestressing and the pylons are integral with
the deck, the pylons may be used during launching to deviate temporary stay cables. This solution
was adopted in the construction of the Millau Bridge in France and the Coast Meridian Overpass
in Canada (Figure 2.18).

The front cable-stayed system includes three main elements (AFGC, 1999):

g A steel portal frame comprising two braced legs supported onto the deck over the webs.
The legs are equipped with base plates embedded with high-strength grout for uniform load
transfer. The legs lodge long-stroke hydraulic cylinders that are used to stress and release
the stay cables. The internal friction of the support cylinders is measured prior to their use
to ensure tight correspondence with the design load procedure for the stay cables. Load
cells may be used for redundant control and real-time recording of the axial load in the
mast, and strain gauges may be applied to the stay cables.
g The stay cables, which may be continuous over steel saddles on top of the mast, or
anchored to the mast by crossing.
g The anchor frames of the stay cables, which typically consist of rectangular frames
embedded onto the deck over the webs with high-strength grout. The vertical prestressing
bars that anchor the frame to the deck are designed for the vertical load component and for
frictional transfer of the longitudinal component of the pull in the stay cables, with
resistance factor f = 0.5. Shear keys may be used to increase the shear capacity of the
connection.

In the longitudinal direction, the mast is pinned to the deck to allow base rotation. In the launch
of cable-stayed bridges with pylons integral with the deck, the pull in the stay cables is adjusted
with strand jacks. A sliding deviation saddle may be applied to the top of the pylon to minimise
longitudinal bending during launching and to simplify pull adjustment and the force equilibrium
of the system (Gale, 2011).

Compared with the use of a launch nose only, the action of the stays modies the launch stresses in
the deck. Before landing at the new pier, the stays resist most of the excess of negative bending
resulting from the use of a short nose. Once the pier has been reached, the stays are relieved and
launch continues with the nose only. The actions of stays and the launch nose are staggered over
the launch and do not overlap one another. Therefore, the nose and the front cable-stayed system
can be designed independently.

97

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


Bridge Launching

The length and weight of the nose are determined rst, and the optimum exural stiffness is
dened using the criteria discussed in Section 2.11.1.1. As the short nose cannot control the
end-of-launch moment MEOL B , the front deck region is designed for the latter, and the cable-stayed
system is designed to cover the excess of negative bending in the cantilever conguration.

The load in the mast and the stay cables is determined for different launch positions so as not to
exceed the SLS edge stress limits specied by the design standards in the front deck region. At the
beginning of launch the stays are unstressed, and they are not necessary until negative bending MB
at the lead pier is equal to MEOL
B . In dimensionless form, the critical advance Lcr that can be
reached without stay cables is

  
Lcr qn L n qn Ln 2 q Ln 2
acr = = + 1.268 n + 0.212 (2.41)
L q L q L q L

In the subsequent launch stages with a acr and up to landing the stays keep MB MEOL B . The
pull Fs in the stay cables may be calculated so as to enforce MB = MEOLB throughout the rest of
launch. Setting b as the inclination of the stay cables on the horizontal, one obtains

 
Fs 1 a qn Ln 0.634 qn Ln 2 0.106
= + + (2.42)
qL sin b 2 q L a q L a

The vertical load that the mast applies to the deck increases the positive bending in the rear span.
In every launch phase, therefore, the pull in the stays must be higher than a minimum value that
depends on the negative exural capacity of the deck, but lower than a maximum value that
depends on the positive exural capacity. The diagram of the minimum and maximum pull in the
stays is determined over the progress of launch (Figure 2.62 shows the case of a short launch nose)
and the pull adjustment procedure must fall between these two curves. The stressing procedure is
discontinuous because pull adjustment occurs during the launch stops. If the two curves are
excessively close or intersect, additional launch prestressing is necessary in the front deck region
to increase the exural capacity. Once the pier has been reached, the nose deection is recovered
by jacking, and the stays are released.

Figure 2.62 Pull adjustment in the stay cables (thicker line). (Reproduced with permission from ASCE)
Pull in the stays

Progress of launch

98

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


Bridge launching

Prior to landing, the weight of the front cantilever is resisted by the lead launch bearings, which are
therefore bigger than those for conventional launch operations. Vertical shear is partly transferred
through the deck, and partly applied by the mast to the rear span of the bridge. Load transfer in
the lead support region often requires oversized webs and vertical prestressing, and has caused
cracking issues in the past that suggest the use of a support diaphragm under the mast.

2.11.3 Temporary piers


The rst PC bridges were launched with temporary piers to shorten the launch spans and diminish
the launch stresses. One temporary pier per span reduces the full-span self-weight bending to
25%, and two piers per span reduce it to 11%. However, the effects of thermal gradients do
not depend on the span length, the effects of misaligned launch bearings and geometry irregula-
rities increase on shorter spans, the temporary piers are more exible vertically than the permanent
piers, and the overall reduction in the launch stresses is much smaller than what one might expect
at rst sight.

With the progressive increase in labour costs, temporary piers have found a specialised niche of use
on long spans that are beyond the possibilities of full-span launching. PC box girders can thus
exceed 100 m spans, and the steel girders can be launched on much longer spans. The 342 m spans
of the Millau Viaduct in France were launched using one temporary pier per span. Long-span PC
box girders are good candidates for the use of one temporary pier per span, as the savings in
reinforcement and launch prestressing, along with a shorter and lighter launch nose, may cover
the cost of the temporary piers, their foundations and the additional labour required for their erec-
tion and nal dismantling and during launching. Bridges with irregular span distribution, includ-
ing short spans t for full-span launching and a few longer spans, are also good candidates.

The temporary piers are more slender and exible than the permanent piers because of the lower
design load and their temporary nature. This is particularly true when tall modular steel towers are
used as temporary piers. Inclined framed props from bridge foundations have also been used for
long-span crossings, and these support systems are even more exible (Figure 2.63). As a result,
under the same tributary deck length, the vertical deection htp of a temporary pier is larger than
the vertical deection hpp of a permanent pier. When one temporary pier is used in each span, the
deection differential generates positive bending in the deck section over the temporary pier, and
negative bending in the deck sections over the permanent piers. In a bridge with piers of similar
height and nal spans of constant length L, these moments are equal and opposite in sign, and
the longitudinal axial stress at the upper and lower edge of the deck
24E  
su,l = + 2
htp hpp zu,l (2.43)
L
depends linearly on the deck depth, and does not depend on the moment of inertia of the
cross-section.

The deck is designed for permanent spans that are twice as long as the launch spans, and the high
depth-to-span ratio during launching results in a stiff response to differential settlement of
supports, with high edge stresses that increase the demand for launch prestressing. The launch
bearings on the temporary piers may be shimmed after loading to recover the differential deec-
tion. In the most delicate cases, the launch bearings are placed on hydraulic jacks for real-time
adjustment of the support reactions. Jack bearings control the effects of bearing misplacement and
geometry imperfections, which are also amplied by the stiff deck response, and simplify the

99

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


Bridge Launching

Figure 2.63 Inclined framed prop for long-span crossing. (Photo: Strabag)

release of the residual support reaction at the end of launching. However, jack bearings are more
expensive than the conventional launch bearings and require skilled crews.

The same reasons discourage the use of many temporary piers in each span. Differential deec-
tions, construction tolerances and stiff deck response reduce the actual possibility of achieving
similar support reactions at all of the piers, unless expensive jack bearings are used at most
supports. The use of multiple temporary piers is, therefore, limited to exible long-span cable-
stayed bridges and the launched decks of tied arches.

The rst temporary piers were made of RC and demolished at the end of launch. Some were cast
in-place as the permanent piers, while others were constructed by lling steel caissons or pipes.
Eventually, modular precast towers were used, composed of RC plates located under the deck
webs, braced transversely by steel diagonals, and prestressed vertically (Figure 2.64). Modular
RC towers are reusable, adaptable to different lengths, less exible than the steel towers, less
sensitive to temperature variations, and generally less expensive when shorter than 710 m.
Steel towers are lighter and can be easily shifted from one alignment to the other (Arcangeli,
1990) for the launch of parallel bridges (Figure 2.65). The steel towers, however, are more exible
and have lower thermal inertia, which may result in larger differences in the vertical deections.

100

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


Bridge launching

Figure 2.64 Modular RC temporary pier. (Reproduced with permission from ASCE)

Figure 2.65 Temporary piers shifted from one bridge to the other. (Reproduced with permission from
ASCE)

101

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


Bridge Launching

The modules of steel towers are typically connected with frontal splices. The behaviour of these con-
nections depends on the tension-to-compression ratio. When the column is always compressed and
the contact with the end plate is milled to bear, the end weld and the frontal bolts are subjected to
minimal stresses of pure shear. In case of load reversal, the net tension is low in most cases, the milled
contact is designed for transfer of compression, and the weld and frontal bolts are designed for
tension and prying action. If the contact is not uniform and the weld breaks in compression, however,
the column can detach from the end plate at load reversal. The robustness of this type of connection
depends on the reliability of the quality-control processes implemented during fabrication and on
the tension-to-compression ratio: if tension is high and quality is uncertain, vertical stressed bars
lodged within boxes welded to the columns may be used to transfer tension from column to column,
bypassing the end plates, and the weld and frontal bolts are designed for shear.

When the temporary piers are very exible, the longitudinal launch loads are resisted with steel
trusses (Figure 2.65), horizontal stay cables anchored to the pier caps of the previous permanent
piers, or inclined stay cables anchored to their foundations or to ground anchors. Horizontal stay
cables have also been used to anchor permanent and temporary piers to the launch abutment,
which diminishes the longitudinal load on the piers and also the net thrust reaction applied to the
abutment foundations. Multiple piers are anchored to the abutment individually to avoid cumu-
lative deections. Horizontal stay cables are also used to stabilise the spandrel columns when the
deck is launched over an arch.

Rear diagonal stay cables are prestressed for stiffness reasons, and front antagonist stay cables
anchored to the foundations of the subsequent piers may be necessary to control load imbalance
and to increase the exural capacity of the pier with additional vertical load (Figure 2.63). The pull
in the stay cables is controlled with load cells or inclinometers to balance the load differential
between breakaway and kinetic friction. The pendular temporary pier shown in Figure 2.66 was

Figure 2.66 Pendular temporary pier

102

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


Bridge launching

pinned at the bottom and stayed to both abutments at the top to avoid longitudinal bending in the
foundation (Rosignoli, 1998d). The strands for the stay cables were contained in self-supporting steel
tubes to avoid the risk of a broken strand falling onto the electric lines. Laminated elastomeric bear-
ings were used for base rotation of the pendular pier.

The launch bearings on tall tubular towers have large vertical load deections, and the towers are
therefore shifted backward from the midspan to give uniform peaks of negative bending in the
deck. This position also allows loading the four columns simultaneously during nose realignment.
The use of one launch bearing on each column increases the labour demand during launching but
simplies the design of the pier cap and improves the longitudinal stability of the pier, as the exural
stiffness of the deck shortens the effective length of the pier by preventing pier-cap rotations. Poor
thermal inertia causes sudden thermal expansion and contraction, and when weight reduction is not
a prime objective the steel pipes are often lled with concrete.

The temporary piers are used successfully only under particular circumstances: short piers,
direct foundations, accessible areas, spans longer than 60 m or very variable, or parallel decks
built subsequently so as to shift either the piers or the deck from one side to the other. Other than
in these cases, on 6070 m spans it is often more cost-effective to reduce the deck slenderness to
12 , H/L , 15 and to use a long launch nose, or a short nose in combination with a front
cable-stayed system.

Finally, the temporary piers may solve unusual launch situations. When a permanent pier supports
the deck at the centre of the bottom slab, temporary columns on either side of the pier support the
deck under the webs during launching. When the deck has varying width or plan curvature, a
temporary column placed alongside the permanent pier supports launch bearings on skidding
shoes that follow the transverse displacement of the deck webs. Temporary piers are also used
to support the closure segments of PC decks launched symmetrically from both abutments.

2.12. Launch stresses in the piers


Incremental launching of PC bridges has been made possible by the commercial availability of
PTFE, a low-cost self-lubricating material that can reduce the sliding friction when in contact with
a polished stainless-steel surface. Previously, the weight of PC decks prevented the use of the launch
bearings used for lighter steel girders. The development of sliding bearings based on the steelPTFE
contact permitted the launch of longer and longer PC bridges. The Veitshochheim Bridge
(Figure 2.14) for the ICE high-speed railway in Germany (Flugel, 1987; Leonhardt, 1991; Theiner,
1987; Zilch, 1987), 1260 m long and weighing 43 300 tonnes, was launched from one abutment with
a total thrust of 18.2 MN. Launching of the 498 m long Sart Canal Bridge in Belgium increased the
launched weight record to 65 000 tonnes. This bridge required the use of four synchronised thrust
systems, each with a capacity of 5 MN and an effective stroke of 2 m (Cremer et al., 2003).

The launch loads of steel girders are rarely critical for the piers, because the girder is launched
without the concrete slab. When a PC box girder is launched uphill, the launch loads are more
demanding, as the friction and launch gradient work in the same direction. The load applied to
a pier during launching has three components:

g The vertical component Rv is the support reaction of the continuous beam at the pier. It
can be eccentric from the pier centreline in the longitudinal direction (load eccentricity
along the launch bearings, or eccentric location of the launch bearings relative to the pier)

103

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


Bridge Launching

and in the transverse direction (misaligned launch bearings), generating two corresponding
couples.
g The longitudinal component, proportional to Rv , is due to the sliding friction and the local
gradient of the launch surface. The launch bearings are aligned with the launch surface,
and inclined bearings apply a longitudinal thrust to the pier.
g The transverse component is due to the guide forces, which do not depend on Rv and are
related to the alignment operations for keeping the deck along the desired trajectory. When
the deck is curved in plan, the transverse component also includes the radial reaction to the
thrust force imparted by the launch systems. When the deck soft has cross-fall, the guide
force balances the radial component of the support reaction at the launch bearings, and no
net transverse force is applied to the pier.

The vertical load is larger than the longitudinal load, which is typically larger than the transverse
load. Therefore, the load applied to the piers lies mostly in the vertical cylindrical surface that con-
tains the deck centreline. Starting from the breakaway friction coefcient C 0f of the Neoon pads
over the launch bearings, the friction angle f is such that
tg f = C 0f (2.44)
Setting a as the launch gradient at the pier, Figures 2.67 and 2.68 dene the vector combinations
for a , f and a . f, respectively. The vertical load applied to the pier is the vector OV, which has

Figure 2.67 Pier-cap forces for launch gradient lower than breakaway friction (a , f ). (Reproduced
with permission from ASCE)

Deck

Uphil
l
Down
hill
T2
H2 H1
T
T1

Launch surface


+
<

R2
N
V
R1

104

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


Bridge launching

Figure 2.68 Pier-cap forces for launch gradient higher than breakaway friction (a . f ). (Reproduced
with permission from ASCE)

Deck

Uph
ill
Dow
T2 nhil
l
H2
T1
H
T


Launch surface
+
>

R2
N
R1
V

a component orthogonal to the launch surface (Rosignoli, 2013)


ON = OV cos a (2.45)
and a tangential component
OT = OV sin a (2.46)

Applying the thrust force rotates the resultant force, and launch starts when the inclination
exceeds the equilibrium angle R2OR1 = 2f. Therefore, the limit equilibrium resultants are
cos a
OR1 = OR2 = OV (2.47)
cos f

When launching uphill, rotating the resultant from OV to OR2 requires a thrust force F equal to
F = TT2 = OT2 OT = OV(cos a tg f + sin a) (2.48)

and the longitudinal load applied to the pier is


cos a  
H = OH2 = OV sin f + a (2.49)
cos f

105

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


Bridge Launching

Downhill launching is more complex. If the launch gradient is lower than friction, a thrust force F
must be applied to rotate the resultant from OV to OR1:
F = TT1 = OT1 OT = OV (cos a tg f sin a) (2.50)
If the launch gradient is higher than friction, the deck slides without any intervention, the force F
is directed uphill to brake the deck, and friction is typically disregarded in its calculation (i.e.
f = 0). In both cases, the longitudinal load applied to the pier is
cos a  
H = OH1 = OV sin f a (2.51)
cos f

To conclude, when the deck is launched uphill, the slope adds to friction, the thrust force applied
to the deck and the longitudinal loads applied to the piers are higher, but the movements of
the deck are better controlled. When the deck is launched downhill, the forces are lower but
their control is more difcult, and a smaller friction coefcient after breakaway may change the
direction of the force to be applied.

During the stops of downhill launch for construction of new deck segments, the deck is anchored
to the launch abutment, and thermal contraction causes the deck to move uphill at the piers. Slope
and friction act in the opposite directions during downhill launch, and pier deections are there-
fore small. Thermal deck contraction can cause uphill pier bending and, ultimately, uphill sliding.
Sometimes it is also necessary to pull the deck backward to recover excessive launch displacements
or to release the abutment anchorages. For all these reasons, the piers are often designed for slope
added to friction, regardless of the planned launch direction.

When the exural capacity of the piers is insufcient to resist launch bending, the piers are
strengthened or cable-stayed. Rear stay cables anchored to the foundations of the previous pier
may be enough when the launch loads do not reverse direction during launching. When the rear
stay cables are prestressed for stiffness reasons, front antagonist cables anchored to the foun-
dations of the subsequent pier may be necessary. Temporary stay cables are less expensive than
strengthening or prestressing of tall piers, and their vertical load component increases the exural
capacity of the pier. If the piers are very tall and exible, the use of synchronised launch reactors
can eliminate the longitudinal load at all piers by balancing action and internal reaction.

The longitudinal load applied to the piers may accidentally exceed the theoretical value based on
the friction and slope. A typical and not infrequent case is the upside-down insertion of a Neoon
pad, with the PTFE in contact with the deck and the neoprene in contact with the launch bearing.
The friction coefcient of neoprene sliding on stainless steel is much higher than the friction
coefcient of PTFE. Tall piers often warrant the use of inclinometers connected to emergency
launch switches to avoid overloading in the case of incorrect operations.

The vertical load applied to the pier may be eccentric from the pier centreline. Under the
assumption of uniform stress distribution on the Neoon pads, the vertical load is applied to the
centre of the launch bearings. The launch bearings can then be shifted backward from the pier
centreline to create an eccentricity that reduces the pier-base bending due to the longitudinal load
applied to the cap.

Imperfections in the geometry of the deck and misaligned bearings make the vertical load eccentric
relative to the launch bearings. This eccentricity is often disregarded in the pier design but can

106

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


Bridge launching

cause signicant effects when the launch bearings are long and the pier is narrow. A notional
eccentricity of +200 mm may be used for pier design (AFGC, 1999). Creep of concrete increases
the pier deection generated by the longitudinal load, and the pier may also have a tolerance in
verticality, especially if built with slip forms. The effects of geometric nonlinearity are investigated
using P-delta analysis.

When the launch gradient is low and the launch force is applied at the abutment, the longitudinal
load applied to the piers during launching rarely governs pier design. Service loads such as bearing
friction and longitudinal wind often prevail. The transverse guide force is also smaller than the
effects of lateral wind on structures and trafc in service conditions, and smaller load factors are
typically used for temporary construction stages.

Long bridges and steep uphill launch gradients may require the use of synchronised Eberspacher
launchers at the launch abutment and at some piers (Figure 2.69). Front stay cables or diagonal
braces anchored to the foundation of the subsequent pier are often necessary on tall piers equipped
with friction launchers. Lighter rear antagonist cables may also be necessary for control of the
longitudinal pier deections in the different phases of the launch cycle. Pier deections are
monitored with conventional topographic techniques; tensioned Invar wires from abutment to
abutment may also be used in short bridges.

The experience gained with synchronised Eberspacher launchers led to the development of the
launch reactors. These electro-hydraulic devices are applied to the pier caps to minimise the
longitudinal loads applied to the piers. They work with the same principles as the Eberspacher
launchers, and a couple is therefore applied to the pier cap in relation to the longitudinal displace-
ment of the deck support point. The launch reactors are positioned such that the support sledge is
at pier centreline when the thrust cylinders are at the middle of the effective stroke. The launch
reactors are controlled by PLC and pier inclinometers to make their response instantaneous and
synchronised with the action of the main launchers at the abutment. The PLC network is set so
that the abutment launchers provide master control of operations, and the pier reactors act as slave
devices based on the feedback received from displacement and pressure sensors. The length of the
thrust cylinders is the same at abutment launchers and pier reactors, to lift the deck and reposition
the support sledges at the same time.

The launch reactors may have a conventional Eberspacher design or an inverted design where four
jacks supported on the pier cap lift and lower a top platform equipped with a low-friction sliding
frame. Longitudinal and transverse double-acting cylinders control the position of the sliding
frame for low-friction simultaneous displacements in any horizontal direction (Bian et al.,
2013). In spite of their cost, these devices may be useful on tall temporary piers and for diminishing
the local effects of eccentric support reactions.

The vertical load may be eccentric also in the transverse direction, as vertical bearing misalignment
and geometry imperfections in the deck cause load differences between the two launch bearings.
Transverse load eccentricity is rarely critical in the permanent piers, which are designed for full
service loads. In temporary piers, it can be controlled with jack bearings capable of adjusting
torsion in the deck and transverse bending in the pier (Figure 2.70).

As cantilevers partially xed at the base, the piers are sensitive to couples and horizontal forces
applied to the top. Pier stability during launch is assessed based on the slenderness ratio

107

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


Bridge Launching

Figure 2.69 Pier launchers for steep uphill launching

l = L0 /r, where r is the cross-sectional radius of gyration and L0 is the effective length. The effec-
tive length is twice the pier length for a xed base, and longer as the degree of base restraint
diminishes. When the slenderness ratio is not very high, bending is calculated neglecting the pier
deections, and the cross-sections are checked for combined axial force and bending with the load
ampliers specied by the design standards, which are typically based on Eulers critical load:

EI
PE = p2 (2.52)
L20

Eulers critical load is calculated using a reduced modulus of elasticity to account for the time-
dependent pier deections generated by the long-term application of the longitudinal launch loads.

108

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


Bridge launching

Figure 2.70 Adjustable jack bearing

If the pier is very slender and deformable, bending is determined by accounting for geometric
nonlinearity.

In the transverse plane, lateral guide forces and wind loading rarely warrant nonlinear analysis. The
deck restrains pier-cap rotations about the longitudinal axis, and the degree of restraint depends on
the torsional/distortional stiffness of the deck and on the span length. Open steel girders on long
spans provide minimal rotational restraint, and the piers are analysed as freestanding cantilevers.
PC box girders on shorter spans provide substantial rotational restraint, and the effective pier
length mainly depends on the lateral exibility of the deck. If the pier base is xed, the transverse
effective length varies from 25% (no lateral deck deection) to 50% (deck exible laterally) of the
longitudinal effective length. When the launch guides are used to brace slender temporary piers,
the permanent piers are designed for full wind loads and full launch guide forces. When the
temporary piers are very tall, four launch bearings may be used to restrain the pier-cap rotations
also in the longitudinal plane (Figure 2.65). As the maximum vertical load is reached in the front
deck region, the stiffness of the launch nose also plays a role in the control of the stability of
temporary piers.

The type of column does not greatly affect the transverse stability of a temporary pier. Columns
and bracing create a transverse truss, and if the braces are inclined an angle a from the horizontal
plane the critical load is (Belluzzi, 1988)
 
PE
Pcr = PE / 1 + (2.53)
2Eb Ab sin a cos2 a

where Eb is the modulus of elasticity of the braces and Ab is their cross-sectional area. This
equation is based on the assumption of instability occurring in the elastic eld. If RC plates of
width b, thickness t, elastic modulus Ec and Poissons ratio n are used for the columns, under the
assumptions that cross-bracing is distributed so as to restrain the plates uniformly along their
vertical edges, the height of the temporary pier is not less than three times the width of the RC

109

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


Bridge Launching

plates, and instability occurs in the elastic eld, the critical load is (Petrangeli, 1993)

p2 Ec t3
Pcr =   (2.54)
3 1 n2 b2

Temporary piers comprising steel or concrete-lled pipes are assessed using similar criteria. Design
standards provide instructions to allow for the effects of geometry imperfections in steel struc-
tures. The design provisions are calibrated with the fabrication and assembly tolerances allowed
in the standard, and may therefore vary from standard to standard. Eurocode 3 (BSI, 2005a)
recognises three levels of tolerances: normal tolerances are used for structures loaded statically,
special tolerances are used for structures different from conventional buildings and where fatigue
phenomena prevail, and more restrictive tolerances are used for crane runways.

The effects of geometry imperfections are considered in the global analysis, in the analysis of the
bracing systems and in the design of members. The tolerances in verticality are considered in
the analysis using an equivalent angular deviation from the verticality f = 0.005kcks , where
 
kc = 0.5 + 1/nc 1.0, ks = 0.2 + 1/ns 1.0, nc is the number of columns and ns is the number
of bracing panels in the temporary pier. The equivalent angular deviation is applied in the
longitudinal direction and in the transverse direction, and the effects are not superimposed.
When the temporary pier has four columns, the angular deviation is also applied with opposite
sign to two planes of columns to induce torsional effects.

The bracing systems are also designed for geometry imperfections. According to Eurocode 3 (BSI,
2005a), the alignment tolerances may be analysed using a lateral midspan deection e = 0.002krL,

where L is the span length of the bracing system, kr = 0.2 + 1/nr 1.0 and nr is the number of
braced members. The geometry imperfections may be modelled using equivalent systems of lateral
forces or by modifying the node coordinates in the structural models.

2.13. Launch stress analysis


In a PC continuous beam cast on falsework in one phase, bending and shear at the end of
construction depend on self-weight and prestressing only. The construction stresses of a launched
bridge are much more complex, and include the effects of deck displacement over the launch
bearings, the effects of geometry imperfections, misalignment of launch bearings and thermal
gradients, and the effects of the irreversible deformations acquired by the deck (because of the
time-dependent behaviour of concrete) during stops in the launch for construction of new segments.

Launch stress analysis requires a considerable amount of calculation. Bending and shear are
calculated for a sequence of closely spaced support congurations that simulate the course of
launching. Analysis accounts for geometry imperfections, misaligned launch bearings, bearing
uplift if decompressed, displacements applied by Eberspacher launchers, thermal gradients, and
the progressive application of new segments and new launch tendons to the continuous beam.
The irreversible creep deections (see Section 3.5.4) acquired by the deck during the previous
launch stops are considered as geometry imperfections during the subsequent launches (Rosignoli,
1998b). The design governing values of bending and shear are enveloped for each curing and pre-
stressing condition of the segments, and when the deck lifts from the launch bearings the system
becomes nonlinear and processing the results becomes more complex.

The complexity of analysis grows over the course of design (AFGC, 1999):

110

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


Bridge launching

g Deck cross-sections and launch prestressing are dimensioned using the criteria discussed in
Chapter 3, using the equations for bending and shear of the perfectly xed beam
(Rosignoli, 1998c). The localised load of pier diaphragms, anchor beams, blisters and
tendon deviators is considered in the analysis. The launch nose is dimensioned using the
approach discussed in Section 2.11.1.
g The rst structural model includes different deck cross-sections, section variations and
localised loads. The model is analysed using a limited number of support positions (1015
per span), a uniform long-term value for axial launch prestressing, thermal gradients
through semi-permanent support congurations (launch stoppages for casting of new
segments), geometry imperfections and bearing misalignment via approximate envelope
values, loads applied by the thrust systems, and a 0.5 MPa compression margin from the
SLS tensile stress limit for self-weight bending and launch prestressing to account for the
time-dependent effects. This level of analysis is often sufcient to determine foundation
loads and estimates of quantities.
g The number of support positions is increased, and thermal gradients, launch prestress
(including time-dependent losses, eccentricity from the deck gravity axis, and local
eccentricity when the bottom slab is thicker in the nal pier regions), geometry
imperfections, bearing misalignment and thrust loads are calculated for each support
conguration. The 0.5 MPa compression margin is kept only in the two front spans, which
are more susceptible to the hyperstatic effects of creep deformations of the deck. The
program signals bearing uplift in the curing area for reanalysis. This level of analysis may
be sufcient for typical bridges.
g The program simulates staged construction of the deck, detects bearing uplift and develops
full nonlinear time-dependent analysis. No margins are kept from the SLS stress limits
specied by the design standards. These analyses may be performed by the independent
checker when the designer adopts the previous level of analysis.

Bridges that are hundreds of metres long require enveloping hundreds of bending and shear
diagrams, each calculated by time-dependent staged construction analysis including multiple load
cases. The number of construction stages corresponds to the number of launch operations, and
each stage typically includes seven steps:

1 Application of a new deck segment to the rear end of the model, application of a new set of
launch tendons, and updating the elastic modulus of concrete for all the deck segments.
The model represents changes in cross-section and steps in the deck gravity axis, and
includes the localised load of anchor beams, pier diaphragms, deviation diaphragms and
tendon blisters. The support that the extraction rails of the casting cell provide to the new
deck segment is modelled using nonlinear compression-only elements to account for uplift
in case of negative reactions.
2 Analysis of the irreversible creep deck deections from the beginning of construction to the
end of the stage at issue. Only the effects of self-weight, launch prestressing and the
irreversible deck deections accumulated in the previous launch stops are considered for the
analysis of the time-dependent effects (see Section 3.5.4). Construction methods such as
balanced cantilever construction require superposition of the time-dependent deections
accumulated during segment casting (Rosignoli, 2013). With incremental launching
construction, and assuming ageing of linear viscoelastic material, the new deformed shape
of the deck replaces the previous one for analysis of the subsequent time-dependent effects
(AFGC, 1999).

111

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


Bridge Launching

3 Analysis of multiple closely spaced support congurations for the rst phase of launching,
from the beginning of segment extraction to launch nose landing at the next pier. Load
cases include self-weight, launch prestressing, launch loads (forces and imposed deections),
thermal gradients, geometry imperfections and misplaced launch bearings. The effects are
calculated using the instantaneous modulus of elasticity, and the irreversible creep deections
are simulated by displacing the support points of the continuous beam accordingly.
4 Application of the new temporary or permanent pier to the model and deection recovery
at the tip of the launch nose. This step interrupts the continuity of the analysis process.
5 Analysis of multiple closely spaced support congurations for the second phase of
launching, from deection recovery to completion of segment extraction. The load cases
include the same load cases as step 3.
6 Envelopment of bending and shear from all the above load cases and structural
congurations.
7 Calculation of the semi-permanent bending moment diagram on launch completion and of
the launch prestress losses due to shrinkage, creep and relaxation for analysis of the
irreversible creep deections during the 12 week construction period of the next segment.

On launch completion, the launch nose is removed from the model, and draped integrative
tendons and superimposed dead loads are applied for long-term analysis of the time-dependent
effects. The support reactions are weighed and corrected during deck placement onto the permanent
bearings, which cancels the hyperstatic effects of the creep deections accumulated by the deck
during launching.

Launch stress analysis requires a considerable amount of calculation also for a steel girder.
Analysis accounts for geometry imperfections, misaligned launch bearings, bearing uplift if
decompressed, launch loads, thermal gradient, and the progressive application of new segments
to the continuous beam. Instead of being affected by creep deections, the steel girder has a
non-launchable prole from the very beginning, as the girder is typically a sequence of rectilinear
segments cambered to shape, and the launch of a cambered geometry on aligned launch bearings
generates hyperstatic stresses.

Such ponderous analysis tasks suggest the use of special computer programs that simulate the course
of launch and iteratively recall a continuous beam solver subroutine from the subsequent structural
congurations. Several methods are available for the analysis of a continuous beam. These methods
are mostly equivalent, as they are based on the same differential equations and require the same
number of integrations (Pestel and Leckie, 1963). The practical interest of a method can therefore
derive from the use of the same type of integrals, or from an agile and repetitive algorithm that
limits the risk of mistakes and that can easily be implemented in a computer program.

The transfer-matrix method (Lacroix, 1967) meets both these requirements, and its advantages
are amplied by the repetitive segment geometry of launched bridges. The method uses a mixed
formulation of the element forcedisplacement relationship to transfer information on the state
of stress of the cross-section from one section to another, and produces systems of equations that
are much smaller than those produced by the stiffness method. In the original formulation of the
method, however, as the course of launch progresses the number of redundant variables increases.

The reduced transfer matrix (RTM) method (Rosignoli, 1997b, 1999a, 1999b) overcomes this
limitation and leads to an exact, simple, elegant and economical algorithm for the solution of a

112

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


Bridge launching

continuous beam with minimal computational effort. The order of the matrices in the solving sys-
tem does not depend on the number of elements used to model the continuous beam, and the accu-
racy of results does not depend on the chosen mesh. Starting from the RTM method for rectilinear
beams, Arici and Granata (2007) have developed an extension of the RTM method to circular
beams discussed in Section 2.13.2.

As the continuous beam is solved using closed-form equations, the RTM method is particularly
advantageous for parametric analysis of design alternatives. The effects of basic design parameters
such as exural and torsional stiffness of deck and launch nose, nose length and weight, and the
elastic modulus of concrete can be compared and optimised with minimal computational effort for
both rectilinear and circular bridges.

Most commercial software programs for structural analysis implement the nite-element method,
and using a software program already available offers advantages in terms of cost, learning curve
and availability of post-processing tools. For those programs that allow the use of custom-written
nite elements, a special element with a moving internal node (Xu and Shao, 2012) is formulated
in Section 2.13.3.

2.13.1 RTM method for rectilinear beams


The RTM method is based on the repetitive multiplication of small square matrices of constant
dimensions the terms of which are obtained from closed-form integration of the elastic beam
differential equations. Consider a continuous beam contained in a horizontal xy plane, with the
x-axis along the gravity axis of the beam, any constraints, and constant or varying cross-section,
with a principal axis of inertia contained in the xy plane. In each cross-section, the state of stress
is dened by three forces (axial force P, in-plane lateral shear Vy and vertical shear Vz ) and three
moments (torsion Mx , longitudinal bending My and in-plane transverse bending Mz ), and the
state of deformation is dened by the three corresponding components of the displacement h and
of the rotation w.

Assuming the presence of only vertical external forces and external couples about the in-plane
axes, the six load effects reduce to three (vertical shear V = Vz , longitudinal bending M = My and
primary Saint-Venants torsion T = Mx ) and the internal work results from vertical deections
h = hz , exural rotations w = wy and torsional rotations q = wx of the cross-sections.
Additional internal work results from warping and cross-sectional distortion; these effects may
be signicant in steel U-girders and braced I-girders, and are analysed using approaches specied
by the design standards, and superimposed.

Most launched bridges are rectilinear or moderately curvilinear, and the use of PC box girders
minimises the effects of torsion. In a rectilinear bridge, torsion and bending are uncoupled, the
unknown deformations during launching are only h and w, and the unknown load effects are only
M and V.

In a PC continuous beam, the effects of prestressing may be analysed using systems of equivalent
forces that include distributed vertical loads along the curved portions of the tendon layout and
localised loads at anchorages and tendon deviators (Cestelli-Guidi, 1987). The loads applied to the
anchorages produce axial and shear forces. The axial forces are balanced by axial forces at the
opposite anchorages and by friction losses along the tendons. These self-balanced systems of axial
forces do not introduce hyperstatic effects in a launched bridge, as the launch bearings allow axial

113

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


Bridge Launching

displacements. Therefore, the effects of prestressing may be calculated by superimposing self-


balanced axial compression on the exural and shear effects of the couples and the vertical forces
of the equivalent system.

So, axial force apart, the elastic system of the beam at each cross-section can be described by a state
array:



h


w
{S } = (2.55)

M


V

The relationship between the state arrays of two subsequent sections can be dened by a matrix [T ]
that transfers the denition of the state array from one section to the other, and is therefore called
the transfer matrix. Between the origin, section 0, and section 1 of the beam (the generic beam
section will be dened by a number, and the support section by a capital letter),

{S}1 = [T ]10 {S}0 (2.56)

Between the two ends of the subsequent beam segment,

{S}2 = [T ]21 {S}1 (2.57)

which can be rewritten as

{S}2 = [T ]21 [T ]10 {S}0 (2.58)

In a beam composed of n subsequent segments, the relationship between the state arrays of the two
end sections can be obtained by progressively multiplying the transfer matrices of the single
segments

{S}n = [T ]nn 1 [T ]nn


2 . . . [T ]1 [T ]0 {S}0 = [T ]0 {S}0
1 2 1 n
(2.59)

and can be summarised as one transfer matrix obtained by multiplication

[T ]n0 = [T ]nn 1 [T ]nn


2 . . . [T ]1 [T ]0
1 2 1
(2.60)

The terms of the transfer matrix [T ]10 between the end sections of the oriented 0  1 beam
segment, long L and of constant exural stiffness EI, are obtained from the equilibrium and
compatibility equations of the elastic beam. These equations are solved through four subsequent
integrations. Integrals depend on the mechanical and geometrical properties of the beam segment,
and the integration constants in the fth column also depend on the external loads applied to the
beam segment:
1
1 L L2 L3
h T15 h

2EI 6EI





f L L2

f

0 1 T25
{S }1 = M = EI 2EI M
= [T ]0 {S }0
1
(2.61)


0 0


1 L T35
V


V

1 1 0 0 0 1 T45 1 0
0 0 0 0 1 0

114

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


Bridge launching

The construction of a PC bridge using incremental launching does not involve many load cases,
and it is easy to create a complete library of integrals. For a distributed load p applied uniformly to
the whole beam segment, by integrating one obtains

pL4 pL3 pL2


T15 = T25 = T35 = T45 = pL (2.62)
24EI 6EI 2

For a distributed load the density of which varies linearly from p0 in section 0 to p1 in section 1
(this is the typical case for a varying-depth launching nose),

4p0 + p1 4 3p0 + p1 3 2p0 + p1 2


T15 = L T25 = L T35 = L
120EI 24EI 6
(2.63)
p + p1
T45 = 0 L
2

The generalisation to any type of distributed load is immediate by observing that the terms in the
fth column are the subsequent integrals of the function p(x), divided by EI in the rst two
rows. The cases of distributed loads applied only to portions of the segment or of varying stiffness
may also be solved by integration. In most cases, however, it is simpler to divide the beam segment
into subsegments fully loaded by the distributed load or with constant stiffness, and then to obtain
the transfer matrix by multiplication.

Although the cyclic reversal of bending and shear suggests the use of axial launch prestressing,
draped tendons have been used for the launch of short spans over temporary piers (Gohler and
Pearson, 2000). For a prestressing force F the eccentricity of which varies linearly from e0 in
section 0 to e1 in section 1 (draped tendons, or rectilinear tendons in a deck region where the
location of the cross-section centroid varies), the beam segment is subject to a moment that varies
linearly from M0 = Fe0 to M1 = Fe1 , and it is therefore
   
F 2e0 + e1 2 F e0 + e1
T15 = L T25 = L T35 = 0 T45 = 0 (2.64)
6EI 2EI

The equivalent vertical load of internal parabolic tendons is introduced by dividing the beam
segment into multiple subsegments for which the vertical load is constant, and the transfer matrix
of the beam segment is obtained by multiplication. For a linear thermal gradient DT between a
warmer upper surface and a colder lower surface, setting H as the deck depth and a as the coefcient
of thermal expansion,

a DT 2 a DT
T15 = L T25 = L T35 = 0 T45 = 0 (2.65)
2H H

Concentrated loads or distortions applied to a section at a distance x from the origin 0 of the beam
segment are introduced by means of an innitesimal subsegment. The 0  1 segment is divided
into three subsegments 0  left, left  right and right  1. The central subsegment is limited
by the sections immediately on the left and on the right of the application point of the load, and
the transfer matrix [T ]10 is always obtained by multiplication:

{S}1 = [T ]10 {S}0 = [T ]1right [T ]right


left [T ]0 {S}0
left
(2.66)

115

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


Bridge Launching

with

x2 x3
1 x 0
2EI 6EI

x x2
0 1 0
[T ]left
0 =
EI 2EI (2.67)

0 0 1 x 0


0 0 0 1 0
0 0 0 0 1

1 0 0 0 Dh
0 Dw
1 0 0
right
[T ]left =
0 0 1 0 DM
(2.68)

0 0 0 1 DV
0 0 0 0 1

(L x)2 (L x)3
1 (L x)
2EI

6EI
0


(L x) (L x)2
0 1 0
[T ]right =
1
EI 2EI
(2.69)
0 0 1 (L x) 0


0 0 0 1 0
0 0 0 0 1

These equations are also used when the concentrated loads or distortions are applied to the starter
section or to the end section of the segment. Localised loads DV are used to represent pier
diaphragms, tendon deviators, anchor beams and blisters of PC box girders, and the deviation
forces of external tendons. Unknown vertical loads are also used to represent the support reactions
of xed bearings. Distortions Dw are used to represent angle breaks in the deck gravity axis due to
exural rotations at the construction joints.

Elastic supports are used to represent the support reactions provided by exible temporary piers.
Setting kV as the vertical elastic constant of the support, the left  right transfer matrix becomes

1 0 0 0 0
0 0
1 0 0

[T ]right = 0 0 1 0 0 (2.70)
left

kV 0 0 1 0
0 0 0 0 1

The nal static system of the deck is a continuous beam, and the temporary launch congurations
are shorter continuous beams with a front cantilever and a rear end segment supported on the
extraction rails of the casting cell. In both conditions, the continuous beam is analysed by taking
the support reactions as redundant variables (Figure 2.71). Starting from the rear end of the deck,
section 0, the transfer matrix [T ]A 0 of the rst deck segment between section 0 and section A is

116

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


Bridge launching

Figure 2.71 Static system of the continuous beam. (Reproduced with permission from ASCE)

0 A B C N1 N 1

RA RB RC RN 1 RN

dened by subsequent multiplication:

{S}A = [T ]A
0 {S}0 (2.71)

The transfer matrix [T ]BA of the rst span is dened in the same way:

{S}B = [T ]BA {S}A (2.72)

and contains the unknown support reaction at section A, among the external loads, in the form
DV = RA in Equation 2.68. It is, therefore,

{S}B = [T ]BA [T ]A
0 {S}0 (2.73)

Proceeding in the same way, one oversteps the front support of the deck (section N) and reaches
the tip of the launch nose (section 1). The relationship between the state arrays of the end sections
of the continuous beam is

{S}1 = [T ]1N [T ]N
N 1 . . . [T ]A [T ]0 {S}0 = [T ]0 {S}0
B A 1
(2.74)

in which each span-by-span transfer matrix, except [T ]A 0 , contains an unknown redundant


variable. The transfer matrix [T ]10 contains all the unknown variables of the continuous beam.
Without unknown support reactions, Equation 2.74 would be four linear equations with eight
unknowns, and four additional equations would need to be written by imposing boundary
conditions at the ends of the deck. The presence of the unknown support reactions requires writing
a corresponding number of additional equations. These may be obtained by calculating the state
arrays of the support sections and by imposing zero deection, or a specied displacement in case
of misplacement of the launch bearings. The number of equations increases during launch due to
the new supports progressively reached by the deck, the solution requires an iterative algorithm,
and the method as a whole loses simplicity and repetitiveness.

The redundant variables are related to the support sections, where they represent the only possible
discontinuity. The transfer matrix method may, therefore, be modied to operate only on the con-
tinuous terms of the state arrays of the support sections, with the remaining terms being calculated
in a second process. Let us consider two subsequent support sections J and K of the continuous
beam that limit the generic span J  K. If the supports are rigid, hJ = hK = 0. Arbitrary vertical
displacements may be applied to the nodes J and K to account for misaligned launch bearings and
deck lifting for replacement of the Neoon pads and insertion of the permanent bearings on
launch completion, so that

hJ = hJ = constant hK = hK = constant (2.75)

117

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


Bridge Launching

and the transfer matrix between the ends of the span includes only three independent variables
1 T T13 T14 T15
K
h
 12 h


0 1


w
T23 T24 T25 w

{S }K = M =
0 0 1 T34 T35
M (2.76)




T45
V
0 0

0 1 V



1 K 0 0 0 0 1 J 1 J

VJ may be extracted from the rst equation and substituted in the other three to express VK as a
function of bending and rotation in the section J:

T12 T 1  
VK = wJ 13 MJ + hK h
 J T15 + T45 (2.77)
T14 T14 T14

This generates a direct relationship between rotations and bending moments in the two support
sections. A reduced state array {SR} may therefore be dened for the support sections,


w

SR = M (2.78)


1

and a RTM of order three may be dened after elimination of VJ ,

(2.79)

In a continuous beam with n rigid supports, the relationship between the reduced state arrays of
the end support sections is

{SR}N = [R]N
N 1 . . . [R]A {SR}A = RA {SR}A
B N
(2.80)
which are two equations with four unknowns. Now that the bending moments MA and MN are
immediately calculable, the two rotations wA and wN are the only unknown terms, and
Equation 2.79 can be solved in closed form:

1  
wA = MN R22 MA R23 (2.81)
R21

Knowing wA , and therefore the reduced state array {SR}A , rotation and bending moment may be
dened using Equation 2.80 for each support section of the continuous beam, and the shear force
may be dened using Equation 2.77. After dening the state arrays of the support sections, shear,
bending, deection and rotation are determined for every cross-section of the deck by means of its
transfer matrix to the nearest support section. Launch stress envelopes can be generated through-
out the length of the deck, and negative and positive bending can be enveloped at the nosedeck
joint for design of joint prestressing.

118

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


Bridge launching

Figure 2.72 Bending moments in the two launch phases

The RTM method can be easily implemented in spreadsheet programs for the analysis of the
launch stresses. The transfer matrices include exact terms, no numerical integration is necessary
and the solution of the continuous beam is immediate. This is particularly advantageous for
parametric analysis as an alternative to complex sequential nite-element analyses.

The course of launching is repetitive. The program is designed for adding a new deck segment and
analysing the deck launch until the casting cell is cleared for the next segment. The program
sequentially analyses two launch phases for every span. The rst phase starts from the end of the
previous launch and ends when the nose tip reaches the landing pier (black lines in Figure 2.72). At
landing, a new transfer matrix and a new RTM are created for the new span. The second launch
phase starts with deection recovery and ends when the segment has been completely extracted
from the casting cell (grey lines in Figure 2.72). Transfer-matrices and RTM are updated with the
course of launch in relation to location of concentrated loads, equivalent prestressing loads and
the other load components. The two sets of load effects are superimposed to generate the envelope
of launch bending (Figure 2.73).

Figure 2.73 Bending envelope

119

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


Bridge Launching

2.13.2 RTM method for circular beams


Curved box girders with a large radius of plan curvature are often analysed as rectilinear beams.
When the deck has a tight radius, shear and bending combine with torsion, and the state array is
expanded to include primary Saint-Venants torsion T = Mx and torsional rotation q = wx (Arici
and Granata, 2007):

h



w



q

{S } = (2.82)
V








M

T

The terms in the transfer matrix of a segment of circular beam of radius R, length L = Rg, exural
stiffness EI and torsional stiffness GJ are obtained from the equilibrium and compatibility
equations of the elastic beam. The integrals depend on the mechanical and geometrical properties
of the beam segment, and the integration constants also depend on the external loads applied to
the beam segment. The extended transfer matrix can be assembled through four submatrices of
order three and two load vectors of order three in the seventh column, as proposed by Arici and
Granata (2007), or can be processed as a whole:

(2.83)

The transfer matrix method for circular beams follows the same analytical approach as the method
for rectilinear beams. In a circular beam composed of n subsequent segments subject to different
loads or with different mechanical properties, the relationship between the state arrays of the two
end sections is obtained by progressively multiplying the transfer matrices of the single segments,
and is summarised in only one transfer matrix.

120

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


Bridge launching

The integration constants are always obtained in closed form. For a distributed load p applied
uniformly to the whole segment, by integration one obtains (Arici and Granata, 2007)
   
pR4 2 + 2 cos g + g sin g pR4 4 + g2 + 4 cos g + g sin g
T17 = (2.84)
2EI 2GJ
   
pR3 g cos g + sin g pR3 2g + g cos g 3 sin g
T27 = + (2.85)
2EI 2GJ
   
pR3 g sin g + 2 2 cos g pR3 g sin g + 2 2 cos g
T37 = (2.86)
2EI 2GJ

T47 = pRg (2.87)

T57 = pR2(1 cos g) (2.88)

T67 = pR(Rg + sin g) (2.89)

For a distributed load the density of which varies linearly from p0 in section 0 to p1 in section 1,
 !   "  
R4 3 3g cos g + g2 3 sin g p0 3 2g + g cos g 3 sin g p1
T17 =
6gEI
4 ! 3  2  "  
R 2g + 15g cos g + 3 g 5 sin g p0 + g3 12g 3g cos g + 15 sin g p1
(2.90)
6gGJ
!   "  
R3 2 + g2 2 cos g 2g sin g p0 + 2 cos g 2 + g sin g p1
T27 =
2gEI
# !     "2   $
R 2 g cos g/2 2 sin g/2 p0 + g2 4 + 4 cos g + g sin g p1
3

+ (2.91)
2gGJ
!   "  
R3 (GJ + EI ) 3g cos g + g2 3 sin g p0 2g + g cos g 3 sin g p1
T37 = (2.92)
2gGJEI
Rg  
T47 = p0 + p1 (2.93)
2
R2 !    "
T57 = g cos g sin g p0 + sin g g p1 (2.94)
g
R 2 !    "
T67 = 2 + g2 2 cos g 2g sin g p0 + g2 2 + 2 cos g p1 (2.95)
2g

The cases of distributed loads applied only to portions of the segment or of varying exural or
torsional stiffness could also be solved by integration. In most cases it is simpler to divide the
segment into subsegments fully loaded or of constant stiffness, to obtain the transfer matrix by
multiplication. For a linear thermal gradient DT between a warmer upper surface and a colder
lower surface, setting H as the deck depth and a as the coefcient of thermal expansion,

R2 a DT   Ra DT sin g Ra DT  
T17 = cos g 1 T27 = T37 = cos g 1
H H H (2.96)
T47 = T57 = T67 = 0

121

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


Bridge Launching

Concentrated loads or distortions applied within the segment are introduced by means of an
innitesimal subsegment, and the related transfer matrix is expressed in analogy with Equation 2.68:

1 0 0 0 0 0 Dh

0 1 0 0 0 0 Dw

0 Dq
0 1 0 0 0

[T ]left =
right
0 0 0 1 0 0 DV
(2.97)

0 0 0 0 1 0 DM

0 DT
0 0 0 0 1
0 0 0 0 0 0 1

When the segment is subject to a support reaction coming from an elastic support, setting kV as the
vertical stiffness and kT as the torsional stiffness of the support, the transfer matrix becomes

1 0 0 0 0 0 0

0 1 0 0 0 0 0

0 0
0 1 0 0 0

[T ]left =
right
kV 0 0 1 0 0 0 (2.98)

0 0 0 0 1 0 0

0 0
0 kT 0 0 1
0 0 0 0 0 0 1

The circular beam is solved using the same approach discussed for the rectilinear beam. The relation-
ship between the state arrays of the end deck sections may be obtained by progressive multiplication
of span-by-span transfer matrices, and additional equations may be written to enforce deection
and torsional rotation compatibility at the support sections. As an alternative, the method may
be modied to operate only on the continuous terms of the state arrays of the support sections,
calculating the remaining terms in a second process. In two subsequent support sections J and K
of the continuous beam, if the launch bearings prevent vertical deections and torsional rotations,
hJ = hK = 0 and qJ = qK = 0. Arbitrary vertical displacements and torsional rotations may be
applied to the supports J and K to account for misaligned launch bearings and deck lifting, so that

hJ = hJ = constant hK = hK = constant qK = qK = constant qK = qK = constant (2.99)

The transfer matrix between the ends of the span now includes only four independent variables:
K


h 1 T12 T13 T14 T15 T16 T17
h





w 0 T22 T23 T24 T25 T26 T27 w








T37 q

q

0 T32 T33 T34 T35 T36





{S }K = V = 0 0 0 1 0 0
T47 V (2.100)


M
0 0
T57 M





0 T54 T55 T56




0 0



T
0 T64 T65 T66 T67
T





1 K 0 0 0 0 0 0 1 J 1 J

122

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


Bridge launching

Extracting VJ and TJ from the rst and third equations, shear and torsion may be expressed as
functions of the bending moment and exural rotation in section J, and of the deections and
torsional rotations of the end sections of the span (Arici and Granata, 2007):
   
T16 T35 + T15 T36 MJ + T16 T32 + T12 T36 wJ
   
+ T17 T36 T16 T37 + T36 hJ h
 K + T13 T36 T16 T33 q K
 J + T16 q
VJ = (2.101)
T16 T34 T14 T36
   
T15 T34 + T14 T35 MJ + T14 T32 T12 T34 wJ
   
T17 T34 + T14 T37 + T34 hK h
 J + T14 T33 T13 T34 q K
 J + T14 q
TJ = (2.102)
T16 T34 T14 T36

Substituting into the second and fth equations in Equation 2.100 leads to a direct relationship
between the exural rotations and the bending moments in the two support sections of the span,
in analogy with Equation 2.79:
K
R11 R12 R13
  
SR K = R21 R22 R23 SR J = [R]K
J SR J (2.103)
0 0 1 J

and to the following terms for the RTM:


   
T26 T14 T32 T12 T34 + T24 T12 T36 T16 T32
R11 = + T22 (2.104)
T16 T34 T14 T36
   
T26 T14 T35 T15 T34 + T24 T15 T36 T16 T35
R12 = + T25 (2.105)
T16 T34 T14 T36
   
T17 T24 T36 T34 T26 + T37 T14 T26 T24 T16
R13 = T27 +
T16 T34 T14 T36
     
T24 T36 T34 T26 h J h  J + T14 T26 T24 T16 T33 q
 K + T13 q K
J q
+
T16 T34 T14 T36
J
+ T23 q (2.106)
   
T54 T12 T36 T16 T32 + T56 T14 T32 T12 T34
R21 = (2.107)
T16 T34 T14 T36
   
T54 T15 T36 T16 T35 + T56 T14 T35 T15 T34
R22 = + T55 (2.108)
T16 T34 T14 T36
   
T17 T54 T36 T34 T56 + T37 T14 T56 T54 T16
R23 = T57 +
T16 T34 T14 T36
     
T54 T36 T34 T56 h J h  J + T14 T56 T54 T16 T33 q
 K + T13 q K
J q
+ (2.109)
T16 T34 T14 T36

123

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


Bridge Launching

In a circular continuous beam with n rigid supports, the relationship between the reduced state
arrays of the end support sections is still given by Equation 2.80, and are two equations with four
unknowns. At the conclusion of matrix assembly, the RTM methods for rectilinear and circular
beams lead to an identical solution. The equations of the RTM method for the rectilinear beam
can actually be obtained as a degenerate case of the RTM method for a circular beam with an
innite radius of plan curvature. The bending moments MA and MN are immediately calculable,
the two rotations wA and wN are the only unknown terms, and the system can be solved for wA
using Equation 2.81, as for a rectilinear beam.

Starting from {SR}A , the reduced state arrays are calculated for all the support sections, the state
arrays of the support sections are completed with shear and torsion by using Equations 2.101 and
2.102, and the state arrays of the deck sections within the spans are obtained by means of their
transfer matrices to the support sections.

Parametric analyses have been performed using the RTM method for circular beams with different
values of the ratio of exural stiffness to primary torsional stiffness, EI/GJ (Granata et al., 2013).
The analysis covered four cases: EI/GJ = 1 for typical PC box girders, EI/GJ = 5 for steel U-
girders launched complete with the concrete slab, EI/GJ = 50 for twin steel I-girders launched
complete with the concrete slab, and EI/GJ = 100 for PC ribbed slabs with a double-T section.
The nose length was 0.3 , Ln /L , 0.8, and the nose weight was 0.03 , qn /q , 0.2. The analysis
showed the following:

g Bending and torsion reach positive and negative peaks in the span behind the launch nose,
like in a rectilinear beam. Bending and torsion increase in the rst phase of launch because
of the increasing length of the front cantilever, diminish at deection recovery, and increase
again in the second phase of launch because of the progressively heavier deck section in the
front span. Deection recovery diminishes torsion signicantly in bridges with a tight plan
radius.
g The peak negative bending in the second phase of launch does not depend much on the
plan radius, especially for low values of EI/GJ, and increases with the increasing EI/GJ.
g Compared with the stress levels in the rear deck region, torsion increases much more than
bending in the front deck region (Figure 2.74). Primary torsion increases with the decrease
in the plan radius and in the EI/GJ ratio.
g For a given plan radius, increasing EI/GJ reduces the peak values of bending and torsion.
For a given value of EI/GJ, shorter plan radii reduce the peak value of bending and
increase the peak value of torsion.
g Launch noses of relative length Ln /L 0.6 are optimal also in curved bridges. The highest
peaks of bending and torsion are reached with short launch noses. Variations in the relative
weight of the nose can only be appreciated in the front span.

The RTM method disregards warping and cross-sectional distortion (Manterola, 2000). Tall piers
and deformable structural systems such as steel U-girders or braced I-girders launched without the
concrete slab are conditions that typically warrant three-dimensional nite-element analysis of the
incremental launching construction of curved bridges.

2.13.3 Finite-element method


Most commercial software programs for structural analysis implement the nite-element method
because of its versatility. Multiple types of elements are available in the same software package and

124

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


Bridge launching

Figure 2.74 Envelope of torsional moment

can be combined in the same structural model, and the models are solved in the same way whatever
the types of element may be. Using the same software package for different problems also offers
advantages in terms of costs and learning curve.

Staged construction of a launched bridge is typically analysed using structural models implement-
ing frame elements. Shell elements are used for buckling analyses of steel girders, performed for
critical support congurations identied using frame-element models, and solid models are used
to investigate stress distributions in specic regions of the deck.

The model is designed according to deck geometry. In a PC deck, special nodes are placed at
the locations of construction joints, cross-section variations, pier diaphragms, and anchor and
deviation points of prestressing tendons. In a steel girder, special nodes are placed at the eld
splices and the changes in plate dimensions, and at the nodes of the bracing systems.
Intermediate joints are then placed within the frame elements to be used as mobile support points
during launch analysis.

Analysis is performed as a staged construction based on short launch steps. In each launch step,
the supports of the previous launch conguration are deactivated, new supports are activated for
the new launch step, and nose deection recovery is enforced in a launch sub-step when the nose
lands at a new pier. A new segment is applied to the rear end of the deck at the end of segment
extraction from the casting cell, the new set of launch tendons is added, and the process is repeated
until deck completion. Finally, the launch nose is removed and the integrative draped tendons are
applied.

Most commercial software programs implement time-dependent formulations of concrete and


prestressing steel, and the time-dependent deections accumulated by the deck during construction

125

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


Bridge Launching

are easy to calculate and to impose as support settlements for analysis of their hyperstatic effects.
Bending, shear and torsion are enveloped from the individual steps of launching, and are post-
processed using spreadsheets outside the analysis platform.

Finite-element analysis of the launch stresses has two weak points: the complexity of parametric
analysis and the complexity of modelling the course of launching. Comparing different design
options requires reanalysis of the model and re-post-processing of the results, while spreadsheet
parametric analyses based on the RTM method are immediate. Modelling the course of launch
also requires a great number of launch phases. A 500 m deck analysed using 2 m launch steps
requires 250 launch steps, and this increases to 300 or more when one considers the application
of new segments and launch tendons and deection recovery. Writing a macro for post-processing
of 300 sets of results can be more demanding than writing a spreadsheet program to implement the
RTM method.

Some commercial software programs allow the use of custom-written nite elements, which are
typically compiled as DDL subroutines and recalled by the software program using proprietary
input/output strings of data. In this case, the use of commercial software can be optimised by using
special frame elements designed to simulate the migration of the support points during launching.
Custom-written nite elements can also be interchanged between different software platforms,
provided that the input/output strings are adapted to the data interchange protocol of the software
package, and the subroutine is recompiled.

The frame element proposed by Xu and Shao (2012) includes an internal support at a varying
location along the element. The position of the support varies over the course of launch and is
dened by its distance 0 x L from one end of the element of length L. The deection is
assumed to be a cubic polynomial on either side of the support point, and the self-weight of the
element is expressed as nodal forces. The element can be connected to other custom elements with
internal supports or to standard frame elements to form bi- or three-dimensional frames. For a bi-
dimensional frame element with internal support, the stiffness matrix is

EA EA
0 0 0 0
L L


(9x + 3L)EI 3(x + L)EI 9EI 3EI

0
xL(L x)
x3 L x2 L xL

 
3 1 3EI EI
+
EI 0
x L L(L x) L

[K ] =

EA
0 0
L


(12L 9x)EI (3x 6L)EI
sym
L(L x)3 L(L x)2


 
3 1
+ EI
Lx L
(2.110)

126

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


Bridge launching

and the mass matrix is

(2.111)

where r is the mass density and A is the cross-sectional area. The terms of the stiffness matrix and
the mass matrix of a three-dimensional element may be found in Xu and Shao (2012).

2.14. Envelopes of bending and shear


Enveloping the bending moment diagrams is denitely quicker and more effective than analysing
hundreds of printout pages. The shear envelopes are prime design tools for the incremental
launching construction of steel girders and prestressed composite box girders with steel
corrugated-plate webs. The envelopes of bending and shear summarise the evolution of the launch
stresses and highlight the deck regions that require strengthening.

The envelopes of launch bending look like the one shown in Figure 2.73. The envelopes typically
include the hyperstatic effects of launch prestressing, as local irregularities in the deck gravity axis
due to changes in the cross-section generate prestress eccentricities that inuence the edge stress
diagrams.

In the rear deck region, the displacement of the pier diaphragms generates regular undulations in
the envelope of launch bending, as the peak positive and negative moments are reached when these
heavier sections are in the midspan. The envelope of the positive moment has its maximum at the
support sections and the envelope of the negative moment has its maximum at the midspan
sections, where the deck is supported when the pier diaphragms are in the midspan.

The undulation in the launch bending envelope depends on the additional weight DQ of the
pier diaphragms with respect to the standard cross-section of the deck. The amplitude of the

127

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


Bridge Launching

undulations is approximately DM h +DQL/8 for both positive and negative bending. As the
additional weight of the pier diaphragms may increase the demand for launch prestressing, the
diaphragms are sometimes cast on launch completion through recesses in the top slab.

Launch prestressing is distributed so as to generate a centroidal resultant force for the midspan
sections. Launch prestressing is often eccentric in the nal support sections of the deck because
a thicker bottom slab lowers the centre of gravity of the cross-section. The peaks in the positive
moment due to prestress eccentricity may be signicant but do not immediately increase the design
edge stresses because of the higher moment of inertia of these cross-sections. Finally, the negative
moment in the front region of the deck is governed by the nosedeck interaction.

Irregular sequences of spans cause anomalies in the stress envelopes. A longer span within a regular
sequence of spans increases the stress envelopes in the whole deck section that passes over that span.
If the longer span is close to the launch abutment, this may cause overdesign of the entire deck. In
these cases, temporary piers are often used to control the stress envelopes, sometimes in combination
with the elimination of the launch bearings on some of the nal piers.

2.15. Design validation by surveying and deck monitoring


In a PC bridge, the launch stresses inuence the design of cross-sections and prestressing systems.
In a steel girder, large patch loads and the torsional stiffness of the long front cantilever may raise
concerns in relation to the stability of the equilibrium (Chacon et al., 2013). Local launch stresses
may be of concern also during the launch of steel trusses (Malite et al., 2000) or cable-stayed grillages
(Gale, 2011). In these cases, the design assumptions may require validation by deck monitoring.

Surveying during construction is typically aimed at achieving a correct nal geometry of the
bridge. In a launched bridge, surveying and deck monitoring during construction are aimed at
assessing the design assumptions and calibrating the structural models (Santos and Min, 2007).
Surveying and deck monitoring provide information about the elastic modulus of concrete, the
coefcient of thermal expansion, the effects of shrinkage and creep, and the local stress dispersal
in the webs over the launch bearings.

Launch surveying requires geometric information on the undeformed deck, and therefore starts
at the beginning of construction. Surveying of a PC bridge involves geometry controls on deck
segments, extraction rails of the casting cell, curing supports, launch bearings, lateral guides, launch
nose and pier verticality. Surveying of a steel girder involves the same type of measurements.

In a PC bridge, six survey front markers are used in every segment. Three markers are placed at a
constant distance from the match-cast joint, and three markers are placed at the same distance
from the rear bulkhead. Four domed-head level control bolts are placed over the webs of the box
girder, where the elevation is not affected by transverse bending or top slab post-tensioning. Two
alignment hairpin markers are placed close to deck centreline, and the centreline of the casting cell
is punch-marked into the hairpins to get the surveying alignment. An additional line of three
markers may be installed at the centre of long segments.

Survey markers are also applied to the top anges of the launch nose. Readings are made prior to
and after match-casting of the front deck segment to ensure the absence of nose displacements.
For surveying of steel bridges, adhesive surveying targets are attached to the surface of the girder,
and level control bolts are drilled into the anges.

128

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


Bridge launching

A theodolite is used to survey plan alignments and elevations relative to the sight to a reference
target. Benchmarks are maintained in the casting facility to check the position of the instrument
and the target and to re-establish the geometry control base if necessary (ASBI, 2008; Rosignoli,
2013). After nishing the new segment, the survey markers are installed, the survey is done and the
data compared with previous readings for the extraction rails to establish the geometry relation-
ships between the extrados and the soft of the segment.

Launch surveying involves data acquisition from deections and boundary conditions, while
deck monitoring involves data acquisition from instrumented sections. The two approaches are
combined for data redundancy. Most reading methods are relatively inexpensive.

Vertical deections of the nose tip and the midspan sections are surveyed in different launch
stages using basic topographic equipment. The deections are read at the deck centreline and over
the webs to include torsional rotations. In multi-girder steel framing systems it is necessary to
undertake a deection survey for every girder, for early detection of distortion and overloading
of cross-frames. When the launch bearings are placed on hydraulic jacks, weighting the support
reactions during launch allows detection and recovery of bearing misalignment (Simon-Talero and
Merino, 2009). The support reactions may also be measured using load cells or strain gauges
embedded in the launch bearing pads, and the nose realignment load may be read as the pressure
from the hydraulic system and related to nose deection.

Deck instrumentation must be planned prior to the start of construction so as not to interfere with
construction operations. In a PC bridge, the sections to be instrumented are typically chosen from
among the most stressed ones (i.e. in the front deck region). This also permits early retrieval of the
rst data. The surfaces monitored by strain gauges overlap, for redundancy of readings and to
eliminate incongruous data (Figure 2.75). Strain gauges located at the bottom of the webs are
oriented in more than one direction in order to measure longitudinal and vertical axial stresses,
as well as tangential stress, when passing over the launch bearings. Strain gauges embedded in
concrete may be combined with external extensometers that measure the distance between two
points of reference bolted to the deck surface. This allows a PC deck to be instrumented after
casting of the segments. When deck monitoring is aimed at investigating the time-dependent
behaviour of concrete, the sections to instrument are chosen from among the most stressed ones
in the nal support conguration. Accelerometers may be applied to external prestressing tendons
to evaluate the tension from their vibration frequencies.

Figure 2.75 Instrumented section. (Reproduced with permission from ASCE)

5 thermocouples

Longitudinal Instrumented block


strain gauges
Strain gauge

3 strain gauges
placed at 45
2 thermocouples

129

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


Bridge Launching

In a PC deck, calculating stresses from recorded strains requires knowing the modulus of elasticity
and the coefcient of thermal expansion of concrete. Embedded strain gauges may be equipped
with thermocouples to identify the thermal strains. Different methods are implemented to measure
the same parameters in order to compare redundant data and eliminate most of the uncertainties.
High local stresses are expected when the instrumented section passes over the launch bearings,
and nite-element analysis may help in dening the theoretical stress distribution and in evaluating
the measured data.

The use of glued strain gauges is common with steel girders (Malite et al., 2000) and steel
corrugated-plate webs. Accelerometers may be applied to the tip of the launch nose to analyse the
vibrations of the front cantilever. Wireless technology may be used to solve access difculties
(Chacon et al., 2009). Several researchers have monitored launched girders and trusses, and have
achieved different levels of accuracy and different amounts of collected data (Lebet, 2004; Malite
et al., 2000; Wipf et al., 2004; Zhang and Luo, 2012).

2.15.1 Modulus of elasticity


Knowledge of the modulus of elasticity of concrete is indispensable for the calculation of stresses
from strains measured using gauges and from surveyed deections. The statistical equations
specied by the design standards to relate the modulus of elasticity to the compressive strength
of concrete are too approximate for analysing the behaviour of instrumented sections, and direct
determination is necessary. Because of the inuence of this parameter on the results of monitoring,
several methods are used at the same time to acquire redundant data.

A rst method consists of measuring the strains of cylindrical specimens subjected to mono-axial
compression. The small dimensions of the specimens and the different curing conditions lead to
approximate data.

A second method is based on tensioning of prestressing tendons, and correlates the average strain
read from many gauges distributed along the cross-section with the force applied by the tendons.
The strains are read immediately before and after each tendon tensioning so as not to superpose
thermal strains and the rst creep strains on the instantaneous elastic ones. The strains are
weighted based on the tributary surface of the extensometers, and their mean value determines the
average deformation plane of the cross-section. The individual strains deviate from this average
plane and, although these deviations do not affect the calculation of the modulus of elasticity, they
indicate that this method tends to become imprecise as the number of tendons increases. Adding
small strains affected by measurement tolerances diminishes the precision of the result, and the
assumption of conservation of plane sections is questionable in these tensioning sequences.

A third method is based on measuring the front support reaction RA in two successive launch
stages (Figure 2.76). It leads to a reliable evaluation of the exural stiffness EI, and on a cloudy
day without signicant thermal gradients it is only affected by the reading tolerances of the jacks
or load cells used for weighting. Setting RA,1 and RA,2 as the measured front support reactions in
the two successive stages of launch, x1 and x2 as the distances between the instrumented section
and the two support points, and z as the distance between the instrumented bre and the cross-
sectional centroid, the exural stiffness is

z  
EI = RA,2 x2 RA,1 x1 (2.112)
12 11

130

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


Bridge launching

Figure 2.76 Evaluation of flexural stiffness. (Reproduced with permission from ASCE)

S R11

B A

x1

S R12

B A
x2

The support reactions are measured in several launch phases to obtain an average value of EI. The
moment of inertia of the cross-section, however, is affected by geometry tolerances in the formed
surfaces and the uncertain tributary width of the slabs.

Other approaches are possible. On a cloudy day, the nose deection may be surveyed on successive
launch phases. The nose deection at landing may be related to the realignment force, and a linear
relationship will conrm the correct structural response of the nosedeck connection. When
Eberspacher launchers are used to thrust the deck, deck lifting may be related to the load applied
by the support jacks. When these methods are combined, nose surveying provides information on
the variation in the elastic modulus of concrete over time, and the launchers provide average data
on young concrete.

Only specimens permit direct measurement of the modulus of elasticity of concrete. In all other
cases, indirect determination is affected by uncertainties in the cross-section geometry. The
modulus of elasticity of a steel girder is typically measured using standard tensile tests on
specimens taken from the girder plates during fabrication.

2.15.2 Correction of results for thermal effects


Correction of the results of deck monitoring for thermal effects is not a simple matter, as the
evolution of temperature in the deck is sharp and can affect the measured strains. It may be
approached in two ways.

In the rst, strains are measured frequently during operations that take a long time (e.g. tendon
tensioning) and the data so obtained are interpolated. Once the elastic deformation has been
estimated, it is possible to cancel the effects of all time-dependent phenomena.

The second approach consists of measuring temperatures in the instrumented section and calcu-
lating the mean thermal strain 10 and the rotation w, which are used to determine the self-balanced

131

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


Bridge Launching

thermal stresses, from which the related strains are subtracted. This method assumes a constant
distribution of temperature throughout the length of the bridge and that the coefcient of thermal
expansion and the elastic modulus of concrete are known.

2.15.3 Coefficient of thermal expansion


Knowledge of the coefcient of thermal expansion of concrete is fundamental when taking
readings over an extended period of time. It is also necessary for determining the strains due to
creep and shrinkage when the instantaneous strain is known (Nam Shiu and Russell, 1983).

The coefcient of thermal expansion depends on the mineralogical nature of the aggregate and the
humidity and temperature of the concrete. It is evaluated using the average strains of the instru-
mented section when the temperature changes. Readings require a sunny day with high thermal
differences. As the coefcient of thermal expansion governs both the mean expansion 10 and the
rotation w of the cross-section, the average strain of the centroidal bre is often used to reduce the
inuence of gradients as much as possible.

2.15.4 Shrinkage
The time-dependent strains recorded by the extensometers combine the effects of shrinkage and of
creep of concrete, and the evaluation of shrinkage requires specic measuring techniques
(Rosignoli, 2000b).

A rst method consists of instrumenting a sample deck segment placed near the bridge to ensure
similar hygrometric conditions. A second method consists of instrumenting a concrete block
placed within the box cell (Figure 2.75). In the rst curing stages, shrinkage is proportional to the
ratio of the drying perimeter to the cross-sectional area. The concrete block is dimensioned with
the same proportions as the deck cross-section, and its ends are waterproofed to reproduce the
hygrometric conditions. As the concrete block is protected within the box cell while the outer deck
surface is exposed to wind and drying, the measured data are approximate. A third method
consists of immersing a strain gauge mechanically isolated from the rest of the section into the
deck concrete, to obtain readings that are independent of the creep strain.

The shrinkage strains have a markedly seasonal evolution, with contraction in the dry periods and
expansion in the humid ones. This contradicts laboratory data, and often produces a total shrink-
age substantially smaller than the expected one. Modern concretes with low water-to-cement
ratios also have less shrinkage than older-generation concretes.

REFERENCES
AASHTO (American Association of State Highway and Transportation Ofcials) (2000) Guide
Specications for Seismic Isolation Design. AASHTO, Washington, DC, USA.
AASHTO (2014) LRFD Bridge Design Specications. AASHTO, Washington, DC, USA.
Abeysinghe R, Gavaise E, Rosignoli M and Tzaveas T (2002) Non-linear pushover analysis of
Greveniotikos Bridge. Journal of Bridge Engineering 7(2): 115126.
AFGC (Association Francaise de Genie Civil) (1999) Guide des Ponts Pousses. Presses de lecole
nationale des ponts et chaussees, Paris, France, p. 240.
Arcangeli A (1990) Viadotto sulla vallata dellOur a Steinebruck, Belgio. LIndustria Italiana
del Cemento, May.
Arici M and Granata MF (2007) Analysis of curved incrementally launched box concrete
bridges using the transfer matrix method. Bridge Structures 3: 165181.

132

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


Bridge launching

ASBI (American Segmental Bridge Institute) (2008) Construction Practices Handbook for
Concrete Segmental and Cable-supported Bridges, 2nd edn. ASBI, Buda, TX, USA.
Belluzzi O (1988) Scienza delle Costruzioni. Zanichelli Editore, Bologna, Italy.
Bennett MV and Taylor AJ (2002) Woronora River Bridge, Sydney. Structural Engineering
International 12(1): 2831.
Bernard J (1997) The Rocher de lEscalade Viaduct, France. Structural Engineering
International (7)1: 1920.
Bian Y, Jiang J, Jing Z, Li A and Liu G (2013) Design and application of hydraulic-walking
incremental launching equipment. Open Construction and Building Technology Journal 7: 17.
BSI (British Standards Institution) (1983) BS 5400-9.1:1983: Steel, concrete and composite
bridges. Bridge bearings. Code of practice for design of bridge bearings. BSI, London, UK.
BSI (2005a) EN 1993-1-1:2005: Eurocode 3 Design of steel structures Part 1-1: General
rules and rules for buildings. BSI, London, UK.
BSI (2005b) BS EN 1998-2:2005: Eurocode 8 Design of structures for earthquake resistance
Part 2: Bridges. BSI, London, UK.
Calvi GM, Priestley MJN and Seible F (1996) Seismic Design and Retrot of Bridges. Wiley,
New York, USA.
Cestelli-Guidi C (1987) Cemento Armato Precompresso. Hoepli editore, Milan, Italy.
Chacon R, Guzman F, Mirambell E, Real E and Onate E (2009) Wireless Sensor networks for
strain monitoring during steel bridges launching. International Journal of Structural Health
Monitoring 8(3): 195205.
Chacon R, Uribe N and Oller S (2013) Numerical validation of the incremental launching
method on a steel bridge through a small-scale experimental study. SEM Experimental
Techniques. http://dx.doi.org/10.1680/bl.59979.133 10.1111/ext.12069.
Chemerinski OI, Seliverstov VA and Zhuravov LN (1996) Launching steel bridges in Russia.
Structural Engineering International (6)3: 7074.
Cremer J, Counasse C and Delforno J (2003) The Sart Canal-Bridge, Houdeng-Aimeries,
Belgium. Structural Engineering International (13)1: 1922.
de Boer J (2011) The benets of jacking and skidding for rapid installation of under- and over-
passes. Structural Engineering International 4: 419425.
De Clercq E and De Ridder J (2003) Curved bridge Mechelse Steenweg, Kortenberg, Belgium.
Structural Engineering International 13(1): 1416.
Dezi L, Menditto G and Rosignoli M (1982) In tema di predimensionamento dei ponti a cassone
in calcestruzzo armato ed in calcestruzzo armato precompresso. LIndustria Italiana del
Cemento 6: 549570.
Favre R, Badoux M, Burdet O and Laurencet P (1999) Design of a curved incrementally
launched bridge. Structural Engineering International (9)2: 128132.
FEM (Federation Europeenne de la Manutention) (1987) FEM 1.001: Regles pour le calcul des
appareils de levage. FEM, Brussels, Belgium.
Flugel K (1987) Der bau der Maintalbrucke Veitshochheim, ein neuer Weltrekord im taktschie-
ben. Tiefbau Berufsgenossenschaft, August.
Fontan AN, Diaz JM, Baldomir A and Hernandez S (2011) Improved optimization formu-
lations for launching nose of incrementally launched prestressed concrete bridges. Journal of
Bridge Engineering 4(2): 461470.
Frizzi D and Giovannini B (1977) Ponte sul Po a Casale Monferrato per l Autostrada dei
Trafori. LIndustria Italiana del Cemento, April.
Gale R (2011) Incremental launching of steel girders in British Columbia two case studies.
Structural Engineering International 21(4): 443449.

133

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


Bridge Launching

Geier R (2010) Design and construction of the Auenbach Bridge, Austria. Structural
Engineering International (20)2: 153156.
Gillet G and Jacquet P (1988) Louvrage 33 sur lautoroute du Littoral a Marseille. Annales
ITBTP, October.
Giovannini B (1972) Un impalcato estruso: il viadotto in Val Restel. LIndustria delle
Costruzioni 32: 314.
Gohler B and Pearson B (2000) Incrementally Launched Bridges Design and Construction.
Ernst, Berlin, Germany.
Granata MF, Margiotta P and Arici M (2013) A parametric study of curved incrementally
launched bridges. Engineering Structures 49: 373384.
Greisch F (1993) Il ponte strallato di Wandre sulla Mosa in Belgio. LIndustria Italiana del
Cemento 63(677): 314333.
Hoeckman W (2001) Bridge over the River Loire in Orleans, France. Structural Engineering
International 11(2): 9498.
Iglesias C (1992) Algunas ideas sobre el predimensionamento de puentes empujados. Hormigon
y Acero 182: 111128.
Jiang F and Yang WWG (1998) Design of the Qi Ao Bridge, China. Structural Engineering
International (8)2: 8182.
Jing T (2004) Bank of China headquarters, Beijing. Structural Engineering International 14(1): 911.
Kelly JM and Naeim F (1999) Design of Seismic Isolated Structures. Wiley, New York, USA.
Kirk M, Tavernier P and Pachetta W (2005) Launching Thurrock Viaduct, United Kingdom.
Structural Engineering International (15)3: 145147.
Lacroix R (1967) La methode des matrices-transfert. Annales ITBTP, MarchApril.
Lebet J (2004) Measurements taken during the launch of the 130 m span Vaux Viaduct.
Steelbridge 2004, Millau, France.
Leonhardt F (1991) Ponte ferroviario sulla valle del Meno a Veitshochheim. LIndustria Italiana
del Cemento, February.
Llombart JA and Revoltos J (1996) Cable-stayed pedestrian bridge, Spain. Structural Engineering
International (6)4: 222223.
Llombart JA and Revoltos J (2000) Petra Tou Romiou Viaduct, Cyprus. Structural Engineering
International (10)4: 233234.
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. IABSE Structural Engineering International (10)3: 193196.
Manterola AJ (2000) Puentes. Colegio de Ingenieros de Caminos, Canales y Puertos, Madrid,
Spain.
Martinez Y Cabrera F and Rosignoli M (2001) Il nodo di Via Palizzi a Milano. LIndustria
Italiana del Cemento, April.
Matildi P and Matildi G (1990) Ponti Metallici, Esperienze Vissute. Cimolai SpA, 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.
Nam Shiu K and Russell HG (1983) Knowledge gained from instrumentation of the
Kishwaukee river bridge. PCI Journal, SeptemberOctober.
Ontario (2006) Incrementally Launched Post-tensioned Concrete Bridge Design, May.
Perez Perez V, Peset Gonzales L, Corres Peireti H and Tarquiz Alfonso F (2011) The launching
of the Pavilion Bridge, Zaragoza, Spain. Structural Engineering International 4: 437442.

134

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


Bridge launching

Pestel EC and Leckie FA (1963) Matrix Methods in Elastomechanics. McGraw Hill, New York,
USA.
Petrangeli MP (1993) Progettazione e Costruzione di Ponti. Masson, Editoriale ESA, Milan, Italy.
Placidi M and Virlogeux M (1991) Il ponte di Trellins: un arco in c.a. sul ume Isere.
LIndustria Italiana del Cemento, October.
Popov OA and Seliverstov VA (1998) Steel bridges on Ankara peripheral motorway. Structural
Engineering International (8)3: 205210.
Rando M, Lomax S and Goberna E (2010) The three Santiago Calatrava bridges in Reggio
Emilia, Italy. Structural Engineering International 20(1): 1820.
Romaro C and Romaro G (2000) Erection of arch and arch-frame bridges. Structural
Engineering International (10)4: 259262.
Rosignoli M (1995) Linterazione avambecco impalcato nei ponti in c.a.p. realizzati per varo
frontale. LIndustria Italiana del Cemento, March.
Rosignoli M (1996) Sul dimensionamento degli impalcati da ponte in c.a.p. realizzati per varo
frontale progressivo. LIndustria Italiana del Cemento 65: 178186.
Rosignoli M (1997a) Inuences of the incremental launching construction method on the sizing
of prestressed concrete bridge decks. Proceedings of the ICE Structures and Buildings
122(3): 316325.
Rosignoli M (1997b) Solution of the continuous beam in launched bridges. Proceedings of the
ICE Structures and Buildings 122(4): 390398.
Rosignoli M (1998a) Launched Bridges. ASCE Press, Reston, VA, USA.
Rosignoli M (1998b) Serio River Bridge, creep and incremental launching. Proceedings of the
ICE Structures and Buildings 128(1): 111.
Rosignoli M (1998c) Nosedeck interaction in launched prestressed concrete bridges. Journal of
Bridge Engineering 3(1): 2127.
Rosignoli M (1998d) Site restrictions challenge bridge design. Concrete International 20(8).
Rosignoli M (1999a) La risoluzione della trave continua nei ponti realizzati per varo frontale.
LIndustria Italiana del Cemento 69(741): 226233.
Rosignoli M (1999b) Reduced transfer matrix method for analysis of launched bridges. ACI
Structural Journal 96(4): 603608.
Rosignoli M (1999c) Presizing of prestressed concrete launched bridges. ACI Structural Journal
96(5): 705710.
Rosignoli M (1999d) Nose optimisation in launched bridges. Proceedings of the ICE
Structures and Buildings 134: 373375.
Rosignoli M (2000a) Thrust and guide devices for launched bridges. Journal of Bridge
Engineering 5(1): 7583.
Rosignoli M (2000b) Creep effects during launch of the Serio River Bridge. Concrete
International 22(3): 5358.
Rosignoli M (2007) Monolithic launch of the Reggiolo overpass. ACI Concrete International 2:
6165.
Rosignoli M (2008) Bridge launching and shifting on the Tanaro River, Alessandria, Italy.
Proceedings of the First ASBI International Symposium on Future Technology for Concrete
Segmental Bridges, San Francisco, CA, USA.
Rosignoli M (2013) Bridge Construction Equipment. ICE Publishing, London, UK.
Rosignoli C and Rosignoli M (2007) Launch and shift of the Tiziano Bridge. ACI Concrete
International 29(10): 4449.
Santos LO and Min X (2007) Load tests of a cable-stayed bridge in Coimbra, Portugal.
Structural Engineering International 17(4): 337341.

135

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


Bridge Launching

Seifried G and Wittfoht H (1979) Die Brucke uber den Shatt-al Arab in Basrah (Iraq). Beton
und Stahlbetonbau 74: 7785.
Simon-Talero JM and Merino RM (2009) Launching of the Vicario Viaduct: Andalucia, Spain.
Structural Engineering International 19(4): 370374.
Tegou SD and Tegos IA (2012) Improvement of the earthquake resistance of incrementally
launched concrete bridges. 15 WCEE, Lisbon, Portugal. http://www.iitk.ac.in/nicee/wcee/
article/WCEE2012_2959.pdf (accessed 20/05/2014).
Theiner J (1987) Weltrekord im Taktschiebeverfahren. Maintalbrucke Veitshoechheim fertiggestellt.
Beton, January.
Virlogeux M, Servant C, Cremer JM, Martin JP and Buonomo M (2005) Millau Viaduct,
France. Structural Engineering International 15(1): 47.
VSL International (1977) The Incremental Launching Method in Prestressed Concrete Bridge
Construction. VSL International Ltd, Koniz, Switzerland.
Wipf T, Phares B, Abendroth R, Chang B and Abraham S (2004) Monitoring of the Launched
Girder Bridge over the Iowa River on US 20. Final Report. CTRE Project 01-108. Center for
Transportation Research and Education, Iowa State University, Ames, IA, USA.
Xu R and Shao B (2012) New beam element for incremental launching of bridges. Journal of
Bridge Engineering 17: 822826.
Zhang Y and Luo R (2012) Patch loading and improved measures of incremental launching of
steel box girders. Journal of Constructional Steel Research 68(1): 1119.
Zhang Z, Zhang J, Hao W, Dai J and Shen Y (2010) Hangzhou Jiandong Bridge designed as a
spatial self-anchored suspension bridge, China. Structural Engineering International 20(3):
303307.
Zhuravov L, Chemerinsky O and Seliverstov V (1996) Launching steel bridges in Russia.
Structural Engineering International 6(3): 183186.
Zilch K (1987) Maintalbrucke Veitshochheim Maschinen und Rusttechnik fur eine ungewohn-
liche Bauaufgabe. BMT Baumaschine und Bautechnik, January.

136

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

Potrebbero piacerti anche