Bridge Launching: Bridge Launching ISBN 978-0-7277-5997-9 ICE Publishing: All Rights Reserved

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 57

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, girder–substringer 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 first step of construction, and a
great number of alternatives for the final structural configuration 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 flexibility 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
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 electrified 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 soffit 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 fixed 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 60–70% of the
main span in order to control overturning during launch. The deck has constant depth, and each

18
Bridge launching

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

deck section is launched on three fixed 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 flexible 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 floorbeams and a concrete deck made
of full-depth precast panels and in-place closure pours over edge girders and floorbeams. 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 fixed 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
Bridge Launching

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

disruption of traffic. 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 field. 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 first 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
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 final
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 first 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 fit 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 final alignment (Figure 2.5). Lightweight concrete was
used for the 168 m main span, and normal concrete for the 128 m anchor span.

21
Bridge Launching

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

When the main span is in the range 30–60 m, the deck is a single-cell, constant-depth PC box
girder. A box girder offers higher structural efficiency 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 60–70 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 4–7 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
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 55–60% of the main span for
varying-depth decks and 65–70% 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 filled 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 configur-
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 traffic-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
traffic 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 16–48 hours
(de Boer, 2011).

Construction of a new bridge by lateral skidding involves casting the deck alongside its final
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 final position of the bridge. When the new deck replaces an

23
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 first 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
Neoflon (neoprene–Teflon) 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 traffic on only
one of the existing bridges, the deck of the second bridge was demolished and replaced with the
new deck. With traffic 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 first deck is launched along the final
alignment of the second deck, and is then shifted laterally to the final 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 first 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 60–70 m, and to steel spans amply exceeding 100 m. Most
incrementally launched PC decks are cast in-place in a fixed 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
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
Bridge Launching

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

PC decks are mostly launched on 40–60 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 50–60 m spans may weigh 1000–1300 tonnes or more. Shipping, site assembly
and final 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 15–20% 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 flexible. Form cambering controls the final span geometry, but load
deflections and residual support action at the application of prestress may cause deck
cracking and steps at the construction joints (Rosignoli, 2013).

26
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
fixity 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
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 40–50 m spans may weigh more than
500–600 tonnes. Like an MSS, a gantry requires shipping, site assembly and final
decommissioning. Setting-up a precasting facility and providing special means of
transportation involve additional investments and financial 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 2–3 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 deflections 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 five 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 Olifant’s River Bridge, which includes 23 PC
railway spans, each 45 m long, to give an overall length of 1035 m. The final structural system
consists of two 11-span continuous beams fixed 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 fixed to the launch abutment. After relieving the
launch prestressing of the first 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 fixed 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 final design of two continuous beams fixed at the abutments, with a
central expansion joint. On launch completion, the upper box girder was permanently fixed to the

28
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 rail–structure 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 configuration used for launching may also be modified 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 800–1000 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
final 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
Bridge Launching

Figure 2.10 Auenbach Bridge. (Photo: Schimetta)

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

30
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 Mainflingen 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 flexibility 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 simplifies rotation about the top of the inclined
legs. The front region of a continuous beam may be made temporarily isostatic by releasing a field
splice in the rear span. The cantilevers of the 147 m main span of the Bayindir Bridge in Turkey were
realigned by releasing field 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 field 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 final 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
configuration is modified to achieve the permanent static system of the bridge, and several alterna-
tives are possible.

g Poly-tetrafluoroethylene (PTFE, Teflon) bearings are used at the piers and the end
abutment, and the launch abutment provides fixity and resists most of the longitudinal
loads. This scheme may be improved with dampers at both abutments and sacrificial 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 fixed 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 fixity 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
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 insufficient 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 flexural 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 flexible 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 deck–arch 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 Isère 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
flexural 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
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 Wümbach Viaduct. (Photo: DB AG)

33
Bridge Launching

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

In the Veitshöchheim 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 filled
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 first spandrel column,

Figure 2.15 Arch stabilisation by means of water filling

34
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, floorbeams 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 flanges of floorbeams, 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 floating cranes or deck-supported derrick
platforms that handle modules of edge girders and floorbeams 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 efficient 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 floorbeams 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 flexural 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 18–20 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 sufficient 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 profile,
and to diminish the drag coefficient.

Cable-stayed box girders are made of steel or PC, or are of composite construction. Streamlined
steel box girders erected with floating 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
Bridge Launching

girders with steel corrugated-plate webs are earning popularity in the 100–200 m span range
because of weight saving, high flexural efficiency, 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 2–3 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) (Löckmann 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
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 final 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 electrified 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 five cable-stayed spans of lengths ranging from 111 to

37
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 profile 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 simplified construction and avoided the use of float-
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
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 first 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 defined 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
Bridge Launching

superimposable onto itself by translation, rotation or rotation–translation. 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 simplifies 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 profile and a law of
cross-fall variation in the area of the bridge. The longitudinal profile 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 soffit, and the cross-section may be distorted to make cross-fall transitions compatible with
the launch requirements.

When the deck carries two-way road traffic, the two halves of the top slab often have opposite
cross-fall, and the deck soffit is a horizontal line orthogonal to the longitudinal plane of symmetry
of the deck. This is the typical configuration also for railway bridges carrying one or two tracks,
with both ballasted track and direct fixation. When the deck carries one-way highway traffic, 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 soffit 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
soffit is modified 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
Bridge launching

1 Rectilinear in plan and with constant longitudinal gradient. The launch lines are rectilinear,
parallel, and on different elevations if the deck soffit has cross-fall. The deck is launched by
pure translation.
2 Rectilinear in plan and circular in profile. 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 profile 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 soffit 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 profile. 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 soffit has cross-fall, the two launch lines are contained in a
flattened 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 profile 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 flexi-
bility of the steel nose simplifies 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 profile, the nose may be set: tangential to
the circle; aligned with the chord (nose tip and nose–deck 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 profile, and lowering with concave profile) 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 significant 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 field splices between segments, but the segments are typically designed to be as long
as possible to minimise field splicing, and the geometry correction is therefore very approximate.
The nose may also be aligned with the chord, so that nose tip and nose–deck 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
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 soffit
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 soffit. 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
Bridge launching

curves should possibly be located outside the bridge. If a transition curve affects the end section of
the bridge, the soffit 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 first 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 field 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 0.5–1.0 m/min. Recovery of the nose deflection at landing
and repositioning of pulling and braking ropes slow down launching, and it may take half a
day to launch a 50–70 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 fit 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 final
structural configuration of the deck.

43
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, 1–4%, is similar to the friction
coefficient of steel–PTFE contact surfaces. Therefore, uphill launching with new or well-greased
Neoflon 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 0.8–1.5 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 first jack reaches its end of stroke, thus avoiding bar

44
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 0.2–0.3%, 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
superfluous 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
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 coefficient 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
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
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 1–50 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 field 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 flow rate of

48
Bridge launching

non-specialised hydraulic pumps (Rosignoli, 2013). In either case, strand jacking is slow compared
with other launch methods, typically in the 2–8 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 specific 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
coefficient 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 coefficient against the deck soffit.
g The deck is lifted, the launch bearings are cleaned from grease, the Neoflon 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 final 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 final 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 2–3% 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 flexural 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
Bridge Launching

In downhill launching with 2–3% gradient, the deck must be pushed to initiate movement, and
then held back when the breakaway friction has been overcome. Two fixed 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 fixed 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 fixed 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 confirm 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
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 soffit. 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
conflicts with the draw cables, tow systems are often the least expensive and most efficient 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 sufficient 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
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
Bridge launching

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

Deck

Reaction beam

based on a friction coefficient 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 field 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 difficulties.

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 field splice, and when the field
splice are bolted numerous holes are available in the web and the bottom flange 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
Bridge Launching

Figure 2.31 Redundant thrust cylinders for a girder–substringer 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 flexural 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 2–3 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 1–2 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 Eberspächer 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
Eberspächer 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
Bridge launching

Figure 2.32 Basic configuration of an Eberspächer 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 coefficient 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
Eberspächer 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 Eberspächer launcher.

55
Bridge Launching

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

1 The support jacks lift the deck 5–10 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 steel–PTFE 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 coefficient 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 flexural 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 coefficient 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
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
influence 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
Bridge Launching

adjacent supports). In very long PC bridges, high thrust forces require a significant 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 influences the sequence of application of launch prestressing. When the first pier is not
very tall, therefore, synchronised pairs of friction launchers may be used at the abutment and
at the first 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 5–10 mm from the reaction blocks during launching. The
upward deflection 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 defines 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 2–3 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 specific operations
in the initial and final 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
sufficient 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 final phases of launch, Rv,L decreases until the end support reaction of the continuous beam
is reached. This end support reaction is often insufficient 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 insufficient for several launches, a second pair of
Eberspächer launchers is applied to a front cable-stayed pier. When the load is insufficient only
in the last 10–30 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
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 flexibility
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
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 1–3% 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 Neoflon 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
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 flow rate
during the return strokes to shorten the cycle time, and the support jacks may be slowed
down during the final 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 Neoflon 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 3–4 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 3–4 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 Eberspächer 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 final phases of launch. These launchers,
however, require deck anchoring during repositioning of the thrust beam when launching along
inclined planes.

61
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 flange
clamps. High contact pressure diminishes the transfer surface of the thrust force, and the field
splices in the bottom flange of the girder are designed to avoid interference with the clamps.
Friction launchers equipped with flange 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
flange clamp.

The clamps may be suspended from the bottom flanges 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 flange clamps, release of wedge brakes, extension of thrust cylin-
ders, application of wedge brakes, release of flange 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 flow 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 flexi-
bility of the deck and the different geometry of the webs. Because of the peculiarities of design, the

62
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
soffit 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 Neoflon 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 soffit 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
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 modification factors that
depend on the importance of the structure and on the classification of the site. The PGA for the
seismic design of the bridge reflects 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) specifies 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 significant, 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 specific 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 flexibility, 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 first 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
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 insufficient, 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, simplifies the pivot bearing, and permits pier design for
the final deck geometry, leaving rotation out of consideration. This scheme is the most
efficient and economical but requires a rigid pier diaphragm that transfers the whole
support reaction to the deck webs during rotation.

In the first two schemes, the rotating support at the pivot pier includes Neoflon 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 Neoflon 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
Neoflon 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 difficult. The deck segment over the rotating support may be cast with the final 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 Neoflon 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 Flèche Bridge in France. Hydraulic jacks with a rotating plunger
were used for the Fontenelle Bridge and the Trith Saint-Lèger 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 Neoflon
plates arranged along a concrete runway beam. This type of support is more flexible, and the
launch movement more regular.

65
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 defines 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 final 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 first 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 final 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 coefficient. 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 coefficient 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–
Teflon contacts. The behaviour of PTFE on polished stainless-steel surfaces is rather variable. The
friction coefficient diminishes from the breakaway friction C 0f,i to the kinetic friction Cf,i during
launching, and both these coefficients depend on several factors:

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

66
Bridge launching

Recommended friction coefficients 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
Neoflon 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 coefficient diminishes with the progress of launching, the breakaway condi-
tion governs the pier design. A breakaway friction coefficient 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 first and, when the front-most bearings start to slide, the friction coefficient 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 influential only in long bridges.

The casting cell typically includes two fixed 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 coefficient may be taken as 10% for Neoflon 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 influential 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 Neoflon 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
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 coefficient 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 Neoflon pads may have a kinetic friction coefficient 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 configurations 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 significantly 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 first condition
is the final 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 flexural launch stresses are only a fraction of
the service stresses, and the deck can be designed for the latter. Longitudinal prestressing is still

68
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 fibre is about one-third of the section
depth, the top fibre modulus is about twice the bottom fibre 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 20–25 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 flexural efficiency and lighter
weight reduce the edge axial stresses of launching and require less launch prestressing. The launch

69
Bridge Launching

tendons are mostly located within the concrete slabs, aligning the prestressing force with the centre
of gravity of the cross-section is difficult with only internal tendons, and some external tendons are
often necessary. This results in more efficient final 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 fluctuations 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 significantly
decreases the weight of the steel webs.

Launching a composite deck complete with the concrete slab would offer significant 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 five 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 flexible, and large
flexural 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 flanges. 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 efficiency rs of the structural design can be
expressed as

p
rs = (2.7)
p+q

70
Bridge launching

Weight reduction limits the influence of the launch stresses and increases the efficiency 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 flexural capacity, and the efficiency 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 efficiency can be expressed as the flexural efficiency rf :

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

The closer rf is to the ideal value of 1, the better the flexural efficiency. 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 fixed 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 fixity 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 final 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 first 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 significant time and costs. Time and costs

71
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 first sight, because of the different
flexibility of the permanent and temporary piers, the construction tolerances and the flexural 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 first-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 deflection. 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, efficient
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 flexural efficiency (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 flexural capacity is so high during
launching that the launch nose may often be avoided or be used only for control of deflections.
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 deflection. 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
web–slab nodes of the deck in order to use the same launch bearings, and the soffit of the nose

72
Next Page
Bridge launching

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

and deck is flush at the joint to enable continuous launching. When the deck launch surfaces are
inclined in the plane of the cross-section, the bottom flanges 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 flexural
and shear stresses in the front region of the continuous beam.

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

73

You might also like