Sei sulla pagina 1di 88

Bridge Launching

ISBN 978-0-7277-5997-9

ICE Publishing: All rights reserved


http://dx.doi.org/10.1680/bl.59979.137

Chapter 3
Prestressed concrete bridges

The most characteristic aspect of the design of a launched bridge is the need to resist the transient
launch stresses. Structures subject to so many and such different load conditions require careful
pre-sizing, to avoid excessive stresses during launching and an excessive cost of the structure in
service. The launch stresses are mostly related to self-weight, and design of the cross-section there-
fore inuences both the load demand and the structural capacity.

Single-cell box girders are preferred for incremental launching construction because of their high
exural efciency. They are efcient at resisting positive and negative bending, as well as torsion
and distortion (Calgaro and Virlogeux, 1994), and forming and launching are less expensive
than with bi- or multi-cellular box girders. Ribbed slabs with a double-T section are used on short
rectilinear spans; they offer easier forming and, when single columns are located under the webs to
avoid pier-caps, the columns may be directly supported on large-diameter drilled shafts to avoid
footings also. Ribbed slabs are also used for incremental launching over arches. Solid and voided
slabs are used for the short spans of highway crossings.

The cross-section is typically designed for launching and adapted to the service stresses and
technological requirements of prestressing. Additional launch prestressing is applied to the front
deck region to withstand higher launch stresses related to the nosedeck interaction. Every cross-
section of the deck is a support section and then a midspan section during launching, and the
cross-section is therefore designed for a bending range:

DM = Mmax Mmin = kgAL2 (3.1)

where g is the specic weight of concrete, A is the cross-sectional area and L is the constant length
of the span. Neglecting thermal gradients, geometry imperfections and misplacement of launch
bearings, with an optimised launch nose the range is 0.16 , k , 0.22 in the deck front region and
0.12 , k , 0.15 in the rear region. For the axial launch prestressing to avoid edge decompression,
the eccentricity generated by DM must keep the prestressing force within the cross-sectional cen-
tral core. The depth hc of the central core depends on the exural efciency rf (Equation 2.10) and
the total depth H of the cross-section:
I
hc = H = rf H (3.2)
zu zl A
and the minimum axial prestressing is therefore
DM gA 2
Fmin = =k L (3.3)
hc rf H
For a given span length, Fmin depends linearly on the weight of the cross-section and inversely
on its depth and exural efciency. The role of the cross-sectional area may be highlighted by

137

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


Bridge Launching

introducing Equation 2.10:

gA2 2
Fmin = kzu zl L (3.4)
IH

As Fmin is a quadratic function of the area, it is imperative to reduce the latter as much as possible.
In order not to penalise the exural efciency, the reduction in the area must be smaller than the
reduction in the moment of inertia, so that the radius of gyration increases. However, structural
and technological constraints prevent excessive lightening of the cross-section, and thus it is often
necessary to increase the moment of inertia as well.

The cross-sectional centroid should be as low as possible to increase the negative exural capacity.
This may be achieved by using a thicker bottom slab, but the cross-sectional area increases and the
structural efciency reduces. Therefore, it is generally convenient to increase the moment of inertia
by increasing the deck depth. The increase in the cross-sectional area is marginal, the moment of
inertia increases in a quadratic ratio, and the radius of gyration increases linearly. The structural
efciency increases due to the higher load capacity, and the exural efciency increases as well.

A deeper deck thus becomes the aesthetic feature of most PC box girders built using full-span
incremental launching. The deck loses the slenderness of a continuous beam, especially on the longest
spans, and assumes the typical look of a simply supported box girder. The dimensional implications
of full-span incremental launching have been studied using statistical analyses (Rosignoli, 1996,
1997a, 1999b). The linear regression of the cross-sectional depth for highway bridges (in metres) is

L
H = 0.94 + (3.5)
22.7

compared with

L
H = 0.04 + (3.6)
21.8

for PC box girders built using other methods (Dezi et al., 1982). The slenderness ratio of highway
bridges is 15 , L/H , 18, with lower values for longer spans, and 12 , L/H , 14 for railway
bridges. When a PC deck has to be more slender, temporary piers may be used to halve the launch
spans (Rosignoli, 2010) as an alternative to more launch prestressing. Steel girders for highway
composite bridges may be launched with slenderness ratios of 25 , L/H , 30.

The aesthetics of a deeper cross-section may be improved by using inclined webs, which shorten
the side wings. The top slab is designed for local bending, punching shear and containment of the
launch tendons. Using the symbols shown in Figure 3.1, the minimum top slab thickness in high-
way bridges is 25 , a/tt , 30, with a lower bound of 0.22 m in the presence of and a lower bound
of 0.26 m in the absence of transverse post-tensioning. Flat ducts may be used for the transverse
tendons to limit the slab thickness. The bottom slab is also made as thin as possible, with a
minimum value of 0.20 m and 0.240.25 m in the regions crossed by the launch tendons; its thick-
ness may be slightly increased in the support regions to improve the negative exural capacity. In
railway bridges, the minimum thickness for both slabs is about 0.30 m.

Web thickness in the midspan sections is governed by the shear demand of full-span launching,
and increases with the span length. For a single-cell highway box girder with internal integrative

138

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


Prestressed concrete bridges

Figure 3.1 Cross-section symbols. (Reproduced with permission from ASCE)

d a d
tt

tw
h
c
tb

prestressing, the thickness of each web may be estimated (in metres) as


 
L
tw = B + 0.02 (3.7)
2100

and with external integrative prestressing

BL 500
tw = + 0.3 0.30.35 m (3.8)
2000

External integrative tendons have become almost standard in modern launched bridges. The
additional web thickness needed to lodge internal parabolic tendons does not provide shear
capacity, as the ducts are empty during launching, and thicker webs increase the shear demand
and reduce the exural efciency of the cross-section. Web thickening in the support regions is
rarely necessary when draped integrative tendons reduce the shear demand on launch completion,
and webs of constant thickness are simpler to reinforce and to form. The webs of single-cell railway
box girders are about 0.50 m thick at midspan for conventional 4050 m spans, and about 0.90
1.00 m thick in the support regions.

Deck depth and web thickness tend to increase the quantities of structural materials in launched
bridges having internal integrative tendons. The average concrete thickness hc (ratio of the total
volume, inclusive of blisters, anchor beams and pier diaphragms, to the top slab surface) of high-
way bridges (in metres) may be obtained as

L
hc = 0.25 + (3.9)
110

while in railway bridges it increases to

L
hc = 0.40 + (3.10)
100

External prestressing and accurate design of the webs, bottom slab, anchor beams, blisters and
pier diaphragms may reduce the total weight considerably. The design of the top slab is more
controversial from an economic perspective, as a thinner slab requires additional reinforcement
or transverse post-tensioning. Prestressing steel (almost always in strand form, as this is easy to
introduce into empty ducts) is 4060 kg/m3 in highway bridges and 6070 kg/m3 in railway
bridges. Reinforcement is 90150 kg/m3 in highway bridges and 140150 kg/m3 in railway bridges.

139

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


Bridge Launching

3.1. Deck segmentation and yard organisation


The very rst launched bridges were based on the idea of positioning precast segments on tem-
porary supports behind the abutment, connecting the segments together with wet joints, and
launching the completed deck into position. Precast segmental construction combined with
launching does offer many advantages.

g Construction of most of the deck is independent of the erection of the piers and abutment
embankment, as the precast segments can be fabricated and stored elsewhere.
g The use of labour and precasting equipment is optimised in a series of repetitive operations,
with excellent quality control and accelerated depreciation of the investment. Precast
segments for launched bridges are absolutely conventional in terms of dimensions and
handling.
g Deck assembly and launching are fast operations that require few workers and little
specialised equipment.
g Launch equipment is easily reusable and relatively inexpensive in terms of the initial
investment and assembly and dismantling costs.

Precast segments can be connected with wet joints, dry joints or epoxy. Epoxy joints are the
standard solution for span-by-span erection of precast segments with self-launching gantries and
for balanced cantilever erection with self-launching gantries or lifting frames. However, they have
two disadvantages in incremental launching applications (Rosignoli, 2013):

g Geometry imperfections increase with the number of joints, and long bridges require
stringent geometry tolerances during short-line match-casting. When the assembly yard
behind the abutment can be as long as an entire span, the segments are positioned on the
support rail and glued to each other, and the new deck section is connected to the rear end
of the launched deck with a wet joint that avoids cumulative errors.
g Launch prestressing is more expensive than for launched bridges with through
reinforcement in the joints, as several design standards specify no edge decompression of
epoxy joints under rare serviceability limit state (SLS) load combinations. Releasing and
repositioning some tendons on launch completion can reduce the prestressing cost, and
results in more efcient permanent prestressing schemes, but these additional operations
take time and increase the demand for skilled labour.

As an alternative, the segments can be precast with reinforcement emerging from the joints and
arranged on the support rail at a distance of 0.20.6 m to splice the rebar and to ll the solutions
of continuity with non-shrinking concrete. Shrinkage cracks in the wet joints are controlled with
local thermal cycles or by casting the joints in the second half of the night to rely on the thermal
expansion of the precast segments produced under daytime temperature (Baur at al., 1966). Site
operations are limited to positioning the segments on the support rail, sealing the joints, applying
prestressing and launching, and the erection rate may be twice the productivity of a cast-in-place
launched bridge.

Most PC launched bridges are cast segmentally behind one or both abutments. The casting yard
includes a casting cell located on a geometric projection of the deck, rebar jigs for partial or total
prefabrication of the steel cage, the thrust systems for the deck, coverings for continued production
in bad weather, prestressing equipment, storage areas distributed under a rail-mounted tower crane
that serves the casting cell, and a batching plant (if ready-mix concrete is not available).

140

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


Prestressed concrete bridges

In most cases, the deck is cast in subsequent segments, and is incrementally launched as soon as
the new segments are completed. The casting cell is xed, and the outer forms are opened prior to
launching and reclosed after segment extraction. The organisation of the casting yard, its length
and its cost depend on multiple factors such as the deck cross-section, launch technology, available
space, total length of the bridge, and length and sequence of the spans.

The length of the bridge, time available for construction and cycle time for segment casting govern
the length of the segments compared with the typical span. The time schedule includes the lead
times for the casting cell, launch nose and launch systems, the assembly duration for the casting
yard, start-up of production, deck construction, nal dismantling of the casting yard, and com-
pletion of the access embankment. Time is less of a constraint when the segments are precast,
as the assembly at site is then a much faster process.

The segment-to-span ratio inuences casting sequences and prestressing schemes, the design of
casting yard and launch equipment, the optimal organisation of production, and the launch
stresses in the deck. The bridge length also inuences the level of industrialisation of the casting
yard. Long bridges comprise a large number of segments and allow higher investments to
mechanise production and reduce the labour demand. Shorter bridges suggest a less expensive
organisation because of the lower number of casting cycles (Rosignoli, 2001).

When the deck is short, it can be cast full-length prior to launching. This is typically the case for
small highway overpasses. The deck can be cast in a full-length formwork or using a mobile casting
cell for long-line match-casting, and the prestressing tendons are anchored to the end diaphragms.
In longer bridges, the deck is launched incrementally, and completion of a new segment is followed
by launching of the entire deck section.

The optimum segment length is dened according to the time schedule and cycle time, and is then
adapted to the typical span length so that the segments correspond to an entire span or a whole
fraction of it. In bridges including a large number of short, equal-length spans, the segments are
often as long as the span to reduce the number of construction joints and to allow reuse of the
same internal formwork. As the span length increases, the segments become one-half or one-third
of the span in order to limit the cost of the installation. In short bridges with long spans it may be
convenient to use much shorter segments, building more segments on a continuous support by
long-line match-casting, and launching them as a whole.

When the number of segments composing the span has been dened, the position of the construc-
tion joints is veried in terms of constructability. The inner tunnel form for one-phase casting of
box girders and ribbed slabs with a double-T section is extracted backwards from the segment on
launch completion, and if the pier diaphragm is cast together with the rest of the section, the joint
is placed immediately in front of the pier diaphragm so that the inner form can be recovered with-
out obstruction (Figure 3.2).

Most segments are cast in two phases (the bottom slab and webs rst, and the top slab second) and
only the form table for the top slab is extracted backwards. The construction joints can then be placed
anywhere, provided that the pier diaphragms are tapered so as not to interfere with the form table
during extraction. The pier diaphragms can also be cast in a second process, through blockouts or
casting pipes in the top slab. Finally, when the launch tendons are coupled at the joints, the couplers
reduce the moment of inertia of the cross-section, which is then placed at the span quarters.

141

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


Bridge Launching

Figure 3.2 Construction joint location for monolithic segments. (Reproduced with permission from
ASCE)

Construction joint Launching

Form extraction

The segments are typically match-cast against the previous ones, and the organisation of the cast-
ing yard therefore hinges on the casting cell. Whatever the subdivision into segments and casting
phases, to build a segment in the shortest time it is necessary to avoid concentrating in the casting
cell activities that might take place elsewhere. Therefore, production is organised into parallel
rather than serial processes, effective handling of semi-worked components is essential, and the
use of a rail-mounted tower crane is decidedly advantageous.

If the bridge is long and enough space is available, it is convenient to create specialised pre-
fabrication areas for the reinforcement cage, tendon ducts and anchor blocks for the launch
tendons. Cage prefabrication outside the casting cell has many advantages. Cage assembly is a
slow and complex operation because of the presence of ducts, anchorages and minor forms.
Prefabrication removes these activities from the critical path, and connes them to a specic area
and segment of organisation. Cage prefabrication improves labour rotation, especially during the
curing stages of concrete. After segment extraction from the casting cell, the outer form is cleaned,
realigned and prepared without constraints. Cage transfer does not compromise the results and
shortens the cycle time. The cage typically includes the rear bulkhead, complicated by the presence
of tendon anchorages, for rapid closure of the casting cell in preparation for the application of the
inner forms.

The rebar jig can be either adjacent to the casting cell or separated from it (Figure 3.3). An adja-
cent rebar jig facilitates towing the cage into the casting cell during deck launching, using rollers
and wheel spacers. A separated rebar jig requires a portal crane or a straddle carrier for cage trans-
fer into the casting cell, but cage transfer is independent of launching, and removal of the rear
bulkhead, joint bush-hammering, and fabrication and tensioning of the launch tendons are
simpler and quicker. The cage can also be fabricated by assembling prefabricated web cages and
slab panels within the casting cell. Rebar jigs for web-cage segments are simple and inexpensive,
and the segments can be handled using the tower crane (Figure 3.4).

Compared with in-place casting on falsework, a higher level of industrialisation and xed logistics
shorten the cycle time, simplify the operations and minimise the labour demand. Parallel activities
maximise productivity, repetitive tasks shorten the learning curve of inexperienced crews, and the
formwork is cleaner and more durable. The level of industrialisation is exible and easy to adapt
to the bridge dimensions and the number of casting cycles. In most cases, however, the setup cost

142

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


Prestressed concrete bridges

Figure 3.3 Rebar jig for the first casting phase; the inner forms ready for transfer can be seen in the
background

Figure 3.4 Prefabricated web cages. (Reproduced with permission from ASCE)

143

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


Bridge Launching

for the casting yard is a signicant portion of the nancial exposure for the project and of the
construction cost of the bridge.

The forming systems may have 26 month lead time, depending on the complexity of deck
geometry. The launch nose and thrust systems have similar lead times if custom designed.
During this time, the casting yard is prepared to receive the casting cell for assembly
(Rosignoli, 2013). Site preparation includes clearing, levelling, grading, casting of foundation
slabs and the foundations for the launch abutment, installation of drainage and storm-water reten-
tion basins, and installation of utilities.

3.2. In-place casting of segments


The casting cycle of a segment is the sequence of operations necessary for its construction. Its
duration depends on the length and complexity of the segment, on cage prefabrication, on the
number of segments to be cast prior to launching, and on the concrete strength requirements of
launch prestressing, and typically varies from 2 working days for 810 m segments to 5 working
days for 2025 m half-span segments. A weekly cycle time is the typical target for half-span seg-
ments so that the concrete can cure over the weekend.

The sequence of operations depends on the casting procedure, on the cage prefabrication, and on
the time required for the concrete to reach the strength necessary to stress the launch tendons.
Natural hardening of a 35 MPa concrete is rarely compatible with weekend curing, and precast
anchor blocks, local thermal cycles with electric wires embedded around the anchorages, or steam
curing of the entire segment are used to shorten the cycle time. In many cases it may be convenient
to increase the design strength of the concrete, as the higher strength allows the strength gain to be
achieved in the scheduled time, lightens the deck, and reduces the inuence of launch stresses,
unexpected events and time-dependent effects.

Whatever the chosen solution, subdividing the deck into segments, subsegments and casting
phases involves the presence of construction joints. Vertical joints are inevitable, and horizontal
joints offer several advantages such as restricting the quantity of concrete processed daily and
reducing the interference of serial operations.

In the vertical joints, the segment is cast against the rear bulkhead of the casting cell, and the
concrete surface is smooth and weakened by the bond breaker. In the new segment, shrinkage and
imperfect adhesion tend to produce horizontal cracks orthogonal to the match-cast joint surface.
These cracks are avoided by bush-hammering or chipping the joint surface (the roughness acts as a
distributed shear key), wetting the joint surface prior to match-casting of the new segment, and
placing a vertical wire mesh near the joint. The use of shear keys is unnecessary, as 10 mm super-
cial irregularities require smaller relative movements for shear transfer than do concentrated
keys. Longitudinal reinforcement crossing the joint further increases the shear capacity of the
connection.

The difference in temperature between the front cured segment and the rear segment in the setting
phase should be reduced until the concrete has reached the strength necessary to activate the
reinforcement and the distributed shear keys, as in these rst curing phases the limited adhesion
is unable to control crack width. In particularly severe climates it may be necessary to maintain
the temperature in the cured segment or to heat its rear end to 15258C, and to cool the setting

144

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


Prestressed concrete bridges

concrete of the new segment. The need for these measures depends mainly on the heat of hydration
of the cement (CEB/FIP, 1994).

Several casting methods are available for the segments: one-phase casting of half-span segments,
two-phase casting of half-span segments in a single or double casting cell, and one- or two-phase
casting of short segments using short- or long-line segmental techniques (Rosignoli, 2013). Several
design and technological aspects of the casting yard organisation are common to the different cast-
ing procedures.

g The casting yard must be easily accessible, levelled and well drained, and its dimensions must be
adequate for setting up the different working areas with minimal interference from one another.
g The foundation beams of the casting cell must be stiff to avoid settlement. Piles, micropiles
or jet grouting may be necessary if there is settling soil. The foundation beams are often
equipped with geometry-adjustment systems for the extraction rails, to facilitate their initial
alignment and their possible realignment after extraction of the rst segment.
g The foundation beams are equipped with lateral guides for the extraction of the rst
segment from the casting cell. The front guide is located at the front end of the casting cell,
and the rear guide is moved progressively forward during segment extraction. After
extraction of the rst segment, only the lateral guide at the front end of the casting cell is
used for the subsequent launches.
g The segment is extracted from the casting cell by sliding it along two full-length
longitudinal rails located under the deck webs. The extraction rails are nished with a
greased, polished, stainless-steel surface, and they work as forms during segment casting
and as low-friction supports during launch. The extraction surfaces are not subject to
exural rotations, and 2 mm poly-tetrauoroethylene (PTFE) plates glued to 1015 mm
steel plates that facilitate detachment from the segment are used for launching. The launch
plates can also be nished with Bakelite instead of PTFE. The launch plates are aligned
along the extraction rails prior to cage fabrication, and they detach from the segment as
they emerge from the lead end of the casting cell during launch. The breakaway friction is
typically 715%, but the weight of the segment is relatively low.

3.2.1 One-phase casting


One-phase casting is the standard solution for solid or voided slabs and for ribbed slabs with a
double-T section. The outer form for solid and voided slabs is simple and inexpensive, and is
typically stripped by lowering using hydraulic wedges. The casting cell for a ribbed slab with a
double-T section is more complex, as the outer forms are rotated away from the segment, and the
inner form must be stripped and collapsed prior to launching (Figure 3.5).

One-phase casting of box girders involves the complex operations of cage fabrication, insertion of
the inner core form, concrete placement and vibration, and removal of the inner form, and the tech-
nological requirements of the different operations often conict with one another. One-phase casting
of long deck segments is used almost exclusively for railway bridges when the owner requires no
horizontal joints in the webs, as the depth of the box girder facilitates access to the inner cell.

The construction schedule is rather rigid because of the interference that occurs at the only
working location and the dead time for concrete curing. Cage prefabrication is often unavoidable,
also in short bridges, and this increases the length of the casting yard and requires expensive cage-
handling equipment.

145

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


Bridge Launching

Figure 3.5 Casting cell for ribbed slabs. (Reproduced with permission from ASCE)

Side form

Tunnel form

Extraction rail Foundation beam

3.2.1.1 Half-span segments


The use of left-in-place inner forms is economically viable only in short bridges, and is typically
allowed only for the voided slabs, as it prevents access to the inner deck surfaces for maintenance
and repair. An inner tunnel form is used in most cases to form the cell of a box girder. The inner
tunnel form remains within the segment during launching (Figure 3.6). After inserting the new
steel cage into the casting cell, the tunnel form is pulled backward from the front segment and
opened within the cage. To avoid geometry conicts that might jeopardise the speed of the
operation, the cage is fabricated on an inner template that stiffens the cage during transfer into
the casting cell. The template is removed backwards on the same rails used for backward extrac-
tion of the tunnel form from the front segment.

The tunnel forms for launched bridges are similar to the tunnel forms used in the movable
scaffolding system (MSS) (Rosignoli, 2013). Operating a long tunnel form is labour intensive due
to the number of articulations and hydraulic systems necessary to collapse the forms, the need to
avoid interference with tendon blisters and deviators and folded form sectors, and the need to
extract the assembly from the segment throughout the rear pier diaphragm. A long tunnel form
is also expensive, although stripping and forming of the inner cell is very fast. The deck segments
may, therefore, be shortened to reduce the investment and to increase the number of casting cycles.

Variations in cross-section geometry, tendon blisters, anchor beams, deviators and pier dia-
phragms govern the modularity of the forming system. Form design may be simplied by using
pier diaphragms of consistent geometry and thickness throughout the length of the bridge, and

146

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


Prestressed concrete bridges

Figure 3.6 Self-launching support truss for the inner tunnel form. (Reproduced with permission from
ASCE)

by widening the opening in the diaphragms as much as possible. The webs and bottom slab may be
thicker at the piers in order to stiffen pier diaphragms with large openings by transverse frame
action.

Tunnel forms carried by full-length trusses are extracted more easily through V-openings in the
pier diaphragms. The truss rolls on saddles supported on the bottom slab of the front segment,
and is received by a second set of saddles supported on the bottom form table of the casting cell.
Independent form modules on adjustable towers require a ush rolling rail for handling by
hydraulic manipulators. The pier diaphragms may also be cast after extracting the inner form,
although this requires a large number of bar couplers and increases the labour demand.

When conventional concrete is used, the top surface of the bottom slab is not formed, and the slab
is cast using concrete having a lower slump in order to avoid reow during web lling. Concrete
pouring starts at the bottom webslab nodes, and the pouring zones are staggered longitudinally
in a sequence designed to avoid cold joints. The use of self-compacting concrete requires a top
form for the bottom slab and air vents to minimise air bubbles, and forms and tie-downs are
designed for full hydrostatic load.

Rened forming systems and heavy lifters for cage transfer involve signicant investment, a cost
that can only be depreciated in long bridges or when the deck is cast on a short aerial casting cell.
Cost-effective one-phase casting solutions may be achieved in short bridges by subdividing the
deck into shorter segments. Independent of the span, and sometimes even of the bridge length,
marked deck segmentation can reduce the construction cost of the bridge (Rosignoli, 2001).

147

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


Bridge Launching

Figure 3.7 Short-line casting and launch. (Reproduced with permission from ASCE)

Thrust cylinders

Fixed casting cell

Curing rail

Three construction methods have been used with short segments: short-line casting and launch,
long-line casting and launch on span completion, and short-line casting and launch on span
completion.

3.2.1.2 Short-line casting and launch


The casting cells for short-line match-casting of launched bridges are conceptually similar to those
used for segmental precasting (Rosignoli, 2013). The geometry-adjustment devices are simpler, no
segment manipulators are needed and only one casting cell is used for the entire bridge. The cast-
ing cell is set on the geometric projection of the deck alignment, at a specied distance from the
abutment (Figure 3.7). The casting cycle provides for the addition of a new match-cast segment
followed by the launch of the entire deck.

Contrary to what happens with the incremental launching of full-length segments, the new seg-
ment is not prestressed. Deck launching extracts the segment from the casting cell and slides it
along a curing rail that progressively receives all the segments of a span. When the planned number
of segments is cast, launch prestressing is applied to the deck section supported on the rail. To be
cost-effective, this casting method requires a long distance between the casting cell and the launch

148

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


Prestressed concrete bridges

abutment. A temporary pier in the end span of the bridge may be used to reduce the exural
stresses in the non-prestressed deck section on the curing rail, and to control overturning in the
initial phases of launch.

A xed casting cell for 5 m segments is designed to produce a segment every day using cage pre-
fabrication. The casting cell may be lengthened to 810 m to produce a segment every 2 days. The
casting cell for a box girder comprises outer steel forms for the rear bulkhead, bottom form table,
webs and side wings, and a movable core form for the box cell, this rolling on runway rails behind
the bulkhead. The casting cell is easy to operate, as the soft width and web depth do not change.
Because of the large number of segments to be cast, the steel forms for the soft, webs and side
wings are typically 67 mm thick. External vibrators may be used to ensure an aesthetic surface
nish in complex segments or congested steel cages.

The core form cantilevers out from a counterweighted tower that rolls longitudinally on runway
rails and carries hydraulic systems for form opening and closure. The core form is inserted into the
casting cell through the rear bulkhead after cage insertion, and is collapsed and extracted back-
ward prior to segment extraction from the casting cell. The core form is a modular assembly of
components that facilitate setting form geometry in the casting of pier diaphragms, anchor blisters
and deviation saddles on a regular basis. Light interchangeable panels allow hand installation in
order to rapidly recongure the casting cell within the cycle time.

The rear bulkhead serves as a template for the ducts of the launch tendons in the slabs and of
the integrative parabolic tendons in the webs. The ducts are secured to holes in the bulkhead, and
the bulkhead must therefore include a hole at every location where a tendon crosses a joint. The
launch tendons are designed to hold the same locations at the joints in order to avoid unnecessary
or overlapping holes in the bulkhead.

The rebar cage and post-tensioning ducts are prefabricated using jigs and templates. One or
multiple rear bulkheads may be used. When the bulkhead is xed, the tendon anchorages are
installed in the casting cell. Two or three rebar jigs may serve the casting cell to allow the cages
to be assembled over a 23 day period. Partial jigs for the bottom slab and webs, top slab and pier
diaphragms may be used to assemble the cage into a nal template prior to its transfer into the
casting cell. The jigs include a stiff framing system that supports the bottom oor, side walls, two
bulkhead templates marked to the proper section size and with locators for the longitudinal ducts,
and side wings and end templates with locators for transverse post-tensioning ducts. Spacers are
provided at the bottom oor, walls and side wings to ensure proper concrete cover. String lines are
used to check the rebar, cover and duct positions from open surfaces.

On cage completion, a lifting truss is connected to the cage with slings and cross-ties. Padded
slings are used to handle epoxy-coated rebars to avoid damaging the coating. Forklifts, straddle
carriers, portal cranes or tower cranes are used to pick up the lifting truss and transfer the cage
into the casting cell. Lifting trusses with scalable anchor points may be required to lift the cage
of deviation segments.

Steel mandrels or inatable hoses are inserted into the ducts throughout the segment to stiffen the
ducts and to prevent demolition in case of grout seepage. The ducts are supported and tied at close
intervals to maintain alignment during concrete pouring. Tight duct support is particularly necess-
ary for the launch tendons in the bottom slab when concrete is placed using transverse ow.

149

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


Bridge Launching

The rst concrete is poured into the central portion of the bottom slab via a chute through the rear
bulkhead or windows in the top slab form. The bottom corners of the webs are lled and
compacted to complete the bottom slab, the webs are lled symmetrically in 0.500.60 m lifts, and
the top slab is poured starting from the centre and side edges and proceeding towards the webs.
The use of retarders in the concrete mix simplies pouring and screed nishing. Mechanical
screeding is followed by a nal check using a straight aluminium edge from the rear bulkhead
to the front segment. Screeds are rarely used for the bottom slab.

The required concrete strength for stripping may be specied at 1520 MPa. The outer form is
stripped by lowering and rotating the web and wing shutters away from the segment, and the core
form is collapsed and extracted backward. Both forms may be handled manually or with hydraulic
systems. Partial application of transverse post-tensioning may be required if the top slab reinforce-
ment is unable to carry the cantilever weight of the side wings; in this case the concrete strength is
often specied at 2530 MPa. Launching often requires a concrete strength of 2530 MPa.

Rear thrust cylinders anchored to reaction blocks behind the rear bulkhead are used to launch the
bridge by pushing the segment out of the casting cell. Modular steel extensions are applied to the
thrust cylinders at the end of each launch cycle, and standard hydraulic systems for push of box
culverts into embankments may be used. Eberspacher launchers at the abutment generate exces-
sive tensile stress in the non-prestressed deck section on the curing rail. Curing rail and sliding
plates are the same as those used in conventional long-segment applications.

The advantages of this construction method are a small, automated casting cell with hydraulic
geometry-adjustment devices, repetitive operations, and efcient labour rotation. The disadvantages
include a long casting yard, the cost of the curing rail, the need for accurate geometry control, the
multiplication of launch operations, and the difculty of handling young non-prestressed segments.

3.2.1.3 Long-line casting and launch on span completion


This construction method avoids frequent launch operations over short distances, which is the
main weak point of short-line casting and launching. The curing rail supports a deck section as
long as an entire span, and a short casting cell shifts along the rail to cast the subsequent segments
(Figure 3.8). The mobile casting cell is similar to the forming systems used for long-line match-
casting of precast segments (Rosignoli, 2013).

When the new deck section is complete, launch prestressing is applied in one operation. Additional
advantages include the application of post-tensioning prior to moving the segments and less
expensive reinforcement, as the cage of an entire span can be fabricated on the curing rail regardless
of the location of the construction joints. When the segments are cast in two stages, the cage of the
rst-phase U-span is fabricated full-length, and the cage of the top slab is assembled segmentally
within the casting cell to simplify progressive removal of the inner forms (Figure 3.9). The disadvan-
tages include the cost of a mobile casting cell and the foundation slab needed for its operation.

The support rail is particularly stiff to prevent settlement due to the segment weight from affecting
the geometry of the casting run. Two longitudinal reinforced-concrete beams located under the
deck webs replicate the geometry of the launch surface, and the deck webs are match-cast over the
support beams. The bottom form table spans the support beams or is supported on shoring
towers, and may be as long as the support rail to save labour and to fabricate a full-length steel
cage during casting of the rst segments, allowing minimal splicing of the rebar and tendon ducts.

150

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


Prestressed concrete bridges

Figure 3.8 Long-line casting and launch on span completion. (Reproduced with permission from ASCE)

Mobile casting cell

Casting
direction
Curing rail

Thrust beam

Draw jack

Short form tables rolling along the support beams have also been used to close the casting cell at
multiple locations. The outer form rolls longitudinally on foundation slabs, and the inner core
form rolls over the bottom form table. The shutters are relatively inexpensive and can accom-
modate different geometries, and the segments can be cast in one or two phases. The shorter the
segments, the larger the number of casting cycles, and the break-even point soon shifts towards
one-phase casting.

The rear segment of the new span is sometimes cast rst. This segment anchors the launch tendons,
and precast anchor blocks are often necessary to shorten the curing time. The standard segments
are cast from the launched deck backward to simplify recovery of the inner form through the pier

151

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


Bridge Launching

Figure 3.9 Full-span fabrication of the U-cage, and segmental top slab cage assembly in the casting cell

diaphragm in the rear segment on span completion. When external integrative tendons or antag-
onist launch tendons require deviators within the box cell, the deviation segments are typically cast
sequentially with the adjacent segments. The deviators may also be cast in a second phase to
assemble the rebar cages through solutions of continuity in the webs.

Combining long-line casting of short segments, incremental launch of entire spans, and side skidding
on launch completion to clear the launch alignment for the construction of a second box girder, may
lead to substantial savings in labour and equipment. Additional savings in prestressing and deck depth
may be achieved with the use of temporary piers on the launch alignment. The 200 m long Tiziano
Bridge in Italy includes two box girders connected at the top slab level and post-tensioned with internal
launch tendons and external integrative tendons. Each 50 m span was built in ve segments using
the 10 m casting cell shown in Figure 3.9 (Rosignoli and Rosignoli, 2007; Rosignoli, 2008).

After application of the integrative tendons, the residual support reactions at the three temporary
piers were relieved by jacking in preparation for skidding (Figure 3.10). The bidirectional launch
saddle used for launching and skidding is shown in Figure 3.11. The launch saddles were placed on
reinforced elastomeric pads to prevent displacement during launching, and Neoon (neoprene
Teon) pads were inserted on launch completion for skidding. After clearing the launch
alignment, the second box girder was incrementally launched with the same casting sequences
used for the rst girder (Rosignoli and Rosignoli, 2008).

3.2.1.4 Short-line casting and launch on span completion


This construction method combines the advantages of the two previous methods and avoids their
weak points (Figure 3.12). A xed casting cell is placed at the rear end of the curing rail. The
segments for the new span are match-cast and progressively pushed along the curing rail. When

152

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


Prestressed concrete bridges

Figure 3.10 Lateral skidding of the first box girder of the Tiziano Bridge

Figure 3.11 Bi-directional saddle for launching and skidding

153

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


Bridge Launching

Figure 3.12 Short-line casting and launch on span completion. (Reproduced with permission from
ASCE)

Auxiliary thrust cylinders Curing rail

Fixed casting cell

Thrust beam

Draw jack

Wet joint

the last segment of the span has reached the necessary strength, the span is pushed forward for wet-
joint connection with the launched deck. After the application of launch prestressing, the deck is
launched to clear the curing rail. If the curing rail is longer than the typical span, new segments can
be cast during deck closure, application of prestressing and launching.

The advantages include a xed casting cell, few launch operations, simple segment extraction from
the casting cell, and avoiding cumulative geometry imperfections with one closure joint in every
span. The disadvantages include the long curing rail, the cost of two thrust systems and the
additional operations for the closure joints.

Short segments introduce additional operations and, as the bridge length increases, this tends
to compromise the savings achieved with a shorter casting cell because of higher labour costs.
Common advantages are the possibility of introducing launch prestressing independent of the
production of segments, and to reduce the cost of prestressing by using tendons as long as an entire

154

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


Prestressed concrete bridges

Figure 3.13 Aerial casting cell on temporary piers. (Reproduced with permission from ASCE)

Working Draw Rear extraction


Thrust beam
platform jack support

RC temporary piers

Form lowering jacks

span. With internal prestressing, short segments can be competitive in short bridges that do not
allow the investment in a 2030 m casting cell. With external prestressing, and especially in the case
of antagonist launch tendons, short segments can be cost-effective also in longer bridges. Short
aerial casting cells have been used successfully in cases where there is soil settlement or no embank-
ment (Figure 3.13).

A xed casting cell for 45 m segments is less expensive than a casting cell for short-line match-
casting of precast segments because of simpler geometry adjustment and the lack of need for
segment manipulators. Productivity is similar to that achievable with half-span casting cells, in spite
of the lower investment. A 2025 m half-span casting cell with weekly cycle leads to 45 m/day
progress. Use of a 45 m casting cell soon leads to a daily cycle being reached, and an 810 m casting
cell with a 2-day cycle and cage prefabrication permits better control of unexpected events.

3.2.2 Two-phase casting in a single casting cell


The most popular casting method for incrementally launched bridges is two-phase casting within
the same formwork, with a horizontal joint at the top of the webs (Figure 3.14). The segment is not
displaced between the two casting operations. The outer form may be designed so that hydraulic
geometry adjustment systems can be used, and thus lower demand for labour, or for the reuse of
modular forms, the costs of which are easier to depreciate over multiple projects.

The most expensive type of outer form consists of steel rib frames hinged at the base and support-
ing steel form panels (Figure 3.15). Being designed for a specic cross-section, this type of form is
rarely reusable. The investment may be worthwhile for long bridges, as in addition to low labour
demand these forms ensure geometry precision and fast opening and closure operations on the
critical path.

Less expensive forming systems consist of braced modular towers supporting wooden form panels.
The forms are stripped by retracting adjustment screws, and the towers may roll on transverse rails

155

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


Bridge Launching

Figure 3.14 Two phase casting in the same formwork

Figure 3.15 Casting cell with rib frames and steel forms. (Reproduced with permission from ASCE)

156

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


Prestressed concrete bridges

for opening and closure of the casting cell. Simplicity and low cost of the forms are balanced by
higher labour costs and slower operations, which limit the use of these forming systems to short
bridges.

An intermediate type of outer form, which is less expensive and more adaptable than the rst, and
faster to use and less labour intensive than the second, involves the use of braced towers installed
on a modular platform supported on hydraulic jacks or wedges. Vertical movements of the
platform facilitate stripping and form setting, the towers can be reused with new cross-section
geometries, the platform can be easily modied to cope with different segment lengths, and the
adjustment screws of the towers are used only for geometry correction.

The geometry of the inner form depends on the location of the horizontal joint. The joint is
typically located at the top of the webs, to hide the joint for aesthetic reasons, but it may also
be located at the bottom of the webs or at mid-depth.

g When the joint is at the top of the webs, a tower crane is used for stripping and handling
the modules of the inner forms. When different full-length sets of inner forms are available
for the pier segments and the midspan segments, the portal crane or straddle carrier used
for cage delivery lifts the inner form in one operation and transports it to a storage
platform located behind the rebar jig (Figure 3.16). Cage prefabrication is limited to the

Figure 3.16 Full-length inner forms for the pier (bottom) and midspan (top) segments

157

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


Bridge Launching

U-section, and the cage for the top slab is fabricated within the casting cell. Modular
design of the forming system facilitates the application of special forms for tendon blisters,
pier diaphragms and anchor beams. After removal of the inner forms, the form table for
the central top slab strip is extracted from the front segment and rolled backwards into
position. The form table may be supported on adjustable towers rolling along the bottom
slab, or on a steel frame spanning extraction rails anchored to the webs. Precast ribs
spanning the webs and supporting galvanised left-in-place forms have been used to avoid
the form table if inspection of the top slab soft is not required (Bennett and Taylor, 2002).
g When the joint is at the bottom of the webs, the inner form is similar to a tunnel form for
one-phase casting. This joint location is rarely used because of the cost, complexity and
labour demand of an inner tunnel form. The tunnel form rolls along the bottom slab
during backward extraction from the front segment, and is supported on the bottom slab of
the new segment during casting of the second-phase pour. The bottom slab resists the
weight of the inner tunnel form and the second-phase concrete, which may be excessive for
2-day curing. The bottom form table for the bottom slab is, therefore, designed for the
entire load.
g Intermediate joints at the centre of the webs are used only when transverse ribs are used to
stiffen the top slab. Mid-depth joints are visible and located in the peak tangential stress
region of the cross-section.

Compared with one-phase casting, scheduling of activities is less restrained, and by prefabricating
the rst-phase cage it is not difcult to achieve weekly casting cycles for the segments. The cycle time
is longer when the cage is assembled in the casting cell. Five working days are typically necessary for
the rst-phase U-segment, and even if the cycle time of the top slab is shorter than 5 days, the total
cycle time often becomes bi-weekly for concrete curing during the weekend. The cycle time does not
depend on the segment length (longer segments just require bigger crews), and the segment length
may be doubled to achieve the same overall productivity. Avoiding the use of a rebar jig and portal
crane for cage delivery compensates for the extra cost of a longer casting cell, and labour savings also
result from the smaller number of launch operations. Two-phase casting in a single formwork is the
standard solution when two adjacent decks are launched simultaneously (Figure 3.17), because this
scheme produces an optimum labour rotation even with cage assembly in the casting cell (Frizzi and
Giovannini, 1977).

3.2.3 Two-phase casting in a double casting cell


Progressive labour specialisation has led to the subdivision of workers operating in a casting yard
into two distinct crews: the ironworkers, often working for a subcontractor, and the carpenters.
When a single casting cell is used, cage prefabrication minimises the interference between the
activities of the two crews. Even so, there is some interference: the adjustment and completion
of the cage after its transfer into the casting cell, the duct splicing, and so on.

Long deck segments may suggest staggering the activities of the two crews longitudinally; for
instance, starting concrete pouring at one end of the casting cell while cage assembly is still going
on at the other end. These solutions are typically a remedy for inadequate scheduling rather than a
means of efcient organisation. In the case of short segments, these solutions often lead to chaotic
situations that are in conict with the cadenced cycles of incremental launching.

A full-length outer form can be combined with a shorter inner form to fabricate the cage and to
cast the segments at different points. The forming system for the Charix Bridge (Figure 3.18)

158

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


Prestressed concrete bridges

Figure 3.17 Parallel launch results in optimum labour rotation. (Reproduced with permission from ASCE)

Figure 3.18 Forming system for the Charix Bridge. (Reproduced with permission from ASCE)

159

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


Bridge Launching

Figure 3.19 Two-phase casting in two casting cells. (Photo: URS)

included an outer form as long as the span, and a shorter inner form that was shifted longitudin-
ally in order to cast short segments using two-phase long-line techniques (Cezard and Servant,
1988).

When the casting yard is long enough, two adjacent casting cells may be used to build the
different elements of the cross-section in operations staggered over time and space. Horizontal
construction joints are placed at the top of the webs, the U-segment is cast in the rear form, deck
launching drags the U-segment into the front form, and the top slab is cast in the front form
(Figure 3.19). This solution offers several advantages. Cage fabrication in one casting cell does
not interfere with the activities of the carpenters in the other casting cell and does not require
heavy handling equipment, the time available for each activity increases, and the critical situations
reduce. The main disadvantage is the higher cost of forming systems and coverings. Another
potential disadvantage is the shrinkage differential between the two elements of the cross-
section, although creep of concrete is high in these rst curing phases and shrinkage tensile stresses
fade quickly.

Unlike two-phase casting in a single casting cell, with a double casting cell the segment is displaced
between the two casting operations. The U-segment may be extracted on xed bearings to avoid
the cost of full-length extraction rails. The casting yard is divided into two zones (Figure 3.20). The
rear casting cell is slightly shorter than the deck segment, and is designed to cast the bottom slab
and webs. The front casting cell is designed to cast the top slab and the remaining portion of
the rst-phase U-segment. The rear U-segment is not match-cast against the front segment, and
the casting cycle is based on the following sequence:

160

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


Prestressed concrete bridges

Figure 3.20 Two-phase casting in two casting cells. (Reproduced with permission from ASCE)

Rear formwork Front formwork


Thrust beam
1st phase casting 2nd phase casting Draw cables

Lifting jacks Completion formwork Abutment


Continuous foundation Launch bearing Draw jacks

1 During curing of the rear U-segment, the inner forms of webs and pier diaphragms are
stripped and the runway rails for the form table of the top slab are applied to the inner
surface of the webs.
2 When the necessary curing has been achieved, the rear U-segment is lifted from the xed
bottom form to insert Neoon pads between the segment and the launch bearings. As the
non-prestressed U-segment is not connected to the front segment, lifting does not cause
stresses in the young concrete.
3 Deck launching extracts the U-segment from the rear casting cell, sliding over the extraction
bearings.
4 On completion of extraction, the form table for the top slab, which is still inside the
previous segment, is pulled backwards and positioned over the U-segment, and the forms
for the side wings and the short closure segment are set up.

Two-phase casting with two casting cells allows the construction of one 2030 m segment per
week. Segment extraction on xed bearings has been gradually replaced by the use of continuous
extraction rails, which, although more expensive, do not require lifting of the rear segment, or the
closure segment, and do not require additional labour for insertion of Neoon pads during
launching.

Two adjacent casting cells can also be used for operations staggered over time but not space. In
each casting cell, the U-segment is cast rst, and the top slab is cast in a second phase. The two
casting cells use one or two sets of inner forms for the webs and pier diaphragms, and one form
table for the top slab, which shuttles back and forth between the two casting cells (Figure 3.21).
The cycle time for the two segments is typically bi-weekly with cage fabrication in the casting cell.
Cage fabrication does not interfere with the activities of the carpenters in the other casting cell,
and does not require heavy handling equipment. The time available for each activity increases, the
number of critical situations is reduced, and bi-weekly launch operations reduce the labour
demand further. The only drawback is the cost of the long casting cell.

3.2.4 Analysis of alternative methods by process simulation


Several alternatives are available for the organisation of the casting yard. The basic decision
between segmental precasting and in-place casting typically leads to in-place casting, unless the
bridge to be launched is part of a large-scale modular project based on precast segmental construc-
tion. Incremental launching construction of precast segmental bridges is very rare nowadays.

161

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


Bridge Launching

Figure 3.21 Twin casting cells. (Photo: URS)

In-place casting offers a large number of alternatives. Three solutions are available for the casting
of short segments (short-line casting and launching, long-line casting and launching on span
completion, and short-line casting and launching on span completion), and three solutions are
also available for longer segments (one-phase casting, two-phase casting in a single casting cell,
and two-phase casting in a double casting cell). A different number of short segments may be used
in one span, and long segments may be one-third of the span, one-half of the span, and even as
long as the span. The steel cage may be prefabricated fully or partially, or may be fabricated within
the casting cell. The launch tendons may cross one, two or three segments, and may be spliced
using couplers or by overlapping. Launch prestressing may be conventional (internal launch
tendons within the slabs), antagonist (a combination of permanent and temporary tendons with
antagonist layout and a centroidal resultant), or a mix of the two. Integrative tendons may also be
external or internal.

A large number of factors inuence the cost-effectiveness of bridge design and project organis-
ation, and the break-even points are different for the same construction method in different
countries, and often even in different regions of the same country. Some factors such as the length
and width of the bridge, the length and uniformity of the spans, span modularity, plan and vertical
curvature, time schedule, and availability and access to the area behind the abutment are related to
the project. Other factors such as logistics and availability and the cost of skilled labour are related
to the project location. The ingenuity of the designer and contractor, the availability of resources,
nancial exposure and the risks of the project, and the contractors propensity to risk are other
determining factors.

Investment in special equipment generates direct costs and nancial exposure; however, the service
life of special equipment is longer than the project duration, and different strategies are available

162

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


Prestressed concrete bridges

to reduce the impact of the investment in the project (Rosignoli, 2013). Leasing the equipment for
the project duration, selling the equipment when it is no longer required, or modifying the equip-
ment and reusing it on new projects are typical strategies. Local scal rules also play an important
role by dening the number of years for full scal depreciation of the investment.

Computer simulations may be used to translate the uncertainties of the project into their potential
impact on project objectives (Shannon, 1992). Project analysis by computer simulation allows
evaluation of the construction processes, prediction of the process performance under different
conditions, analysis of the functional relationships, sensitivity analysis of the process performance,
process optimisation and bottleneck analysis (Marzouk et al., 2007). Computer simulation is a
powerful tool for modelling construction processes and comparing alternatives, although its appli-
cation within the bridge industry is still limited.

Process analysis is based on formulating a simulation model for the process, running the simu-
lation, and analysing and comparing the results. The process may be progressively rened and
optimised during the simulation. General-purpose simulation engines such as AweSim (Pritsker
et al., 1997), GPSS/H (Crain, 1997) and MicroCYCLONE (Halpin and Riggs, 1992) have been
designed for the modelling and analysis of site-level processes which are cyclical in nature.
These programs are used to model construction operations that involve task interaction, and the
resource ow routes through the work tasks are the basic rationale for the sequencing of construc-
tion operations.

The simulation network for the construction of a PC bridge using incremental launching involves
three main operations:

g Preparation of the casting yard. This may include: the design and fabrication of the inner
and outer forms, extraction rails, launch nose and launch systems; the design and
fabrication of the rebar jigs, lifting truss and portal crane for cage delivery; the design and
construction of the foundations and production support facilities; and the shipping,
assembly and commissioning of equipment. In the most complex cases, the casting yard
may also include batching plant, concrete delivery lines and concrete distribution systems.
g Deck fabrication. This includes repetitive casting cycles of deck segments. Each casting
cycle includes the fabrication of the steel cage and prestressing ducts, the handling of the
external and internal forms, pouring of concrete, nishing of the top slab, segment curing,
strand insertion and the application of prestressing, and deck launching. Different deck
casting processes include different numbers of segments, different cycle times and different
resources.
g Deck completion. This includes, at every pier, deck lifting, removal of the launch bearings,
weighing and adjustment of support reactions, installation and grouting of the permanent
bearings, curing of the bearing grout and deck lowering. During bearing replacement, deck
completion also includes the application of integrative prestressing, tendon grouting and
deck nishing.

The simulation network is identied through the tasks involved in each operation. Resources are
assigned to the tasks, and the assignments are progressively calibrated during the optimisation
process. New tasks may be inserted into the process according to need. For example, two-phase
casting in a single casting cell with cage fabrication within the casting cell, internal launch and
integrative tendons, splicing of launch tendons by overlapping, no transverse post-tensioning in

163

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


Bridge Launching

the top slab, and launching with rear thrust beam and draw cables, includes four main operations
and 63 tasks:

g Casting of bottom slab and webs. This operation includes 25 tasks: (1) cleaning, alignment
and oiling of the outer form; (2) cleaning and lifting of the bottom form table; (3) cleaning,
positioning and oiling of the rear bulkhead; (4) greasing of the extraction rails and placement
of the launch plates; (5) oiling of the bottom form table; (6) placement of spacers for the outer
concrete cover; (7) placement of the outer rebar grid for the bottom slab and webs;
(8) placement of the support rebar for the top rebar grid of the bottom slab and for the ducts
of the launch tendons; (9) xing the anchorages of the launch tendons to the rear bulkhead;
(10) placement of tested ducts for the launch tendons in the bottom slab and splicing to the
ducts emerging from the front segment; (11) placement of the top rebar grid of the bottom
slab; (12) placement of tested ducts for the parabolic integrative tendons in the webs, and
splicing to the ducts emerging from the front segment; (13) temporary positioning of
anchorages of the parabolic integrative tendons; (14) placement of the inner rebar grid of the
webs; (15) completion of the rebar cage for the pier diaphragm and tendon blisters;
(16) application of the inner forms for webs, anchor blisters and pier diaphragm; (17) xing
the anchorages of the parabolic integrative tendons to the anchor blister forms; (18) geometry
control of the completed cage; (19) insertion of rubber pipes into the tendon ducts; (20)
pouring and vibration of the concrete; (21) nishing of the top surface of the bottom slab and
of the construction joints at the top of the webs; (22) concrete curing; (23) removal of the inner
forms and rubber pipes from tendon ducts; (24) cleaning, inspection and repair of the interior
surfaces; and (25) application of the support frames for the rolling form table for the top slab.
g Casting of the top slab. This operation includes 16 tasks: (1) rolling the form table from the
front deck segment backwards into position; (2) cleaning, oiling and lifting of the form
table; (3) placement of spacers for concrete cover; (4) placement of the bottom rebar grid;
(5) placement of the support rebar for the top rebar grid and the ducts of the launch
tendons; (6) placement of scuppers and other embedded items that interrupt the top rebar
grid; (7) xing the anchorages of the launch tendons to the rear bulkhead; (8) placement of
tested ducts for the launch tendons, and splicing to the ducts emerging from the front
segment; (9) placement of the top rebar grid; (10) placement of minor embedded items;
(11) geometry control of the completed cage; (12) insertion of rubber pipes into the tendon
ducts; (13) pouring and vibration of the concrete; (14) screed nishing of the top slab
surface; (15) concrete curing; and (16) removal of rubber pipes from the tendon ducts.
g Application of launch prestressing. This operation includes 10 tasks: (1) removal of the rear
bulkhead; (2) stripping of the form table for the top slab and lowering onto the support
rollers; (3) stripping and opening of the outer form; (4) stripping and lowering of the
bottom form table; (5) bush-hammering and pressure washing of the joint surface;
(6) cleaning the tendon anchorages; (7) cleaning the air vents and washing the ducts;
(8) inserting strand into the ducts; (9) application of the anchor plates and wedges; and
(10) stressing the launch tendons.
g Launching. This operation includes 12 tasks: (1) positioning and anchoring the thrust beam
to the rear end of the segment; (2) insertion of the draw cables into the anchorages;
(3) positioning the thrust jacks; (4) distribution of personnel on the pier caps; (5) testing the
launch systems; (6) isotensioning of the draw cables; (7) release of the deck anchor systems;
(8) staged launching; (9) nal application of the deck anchor systems; (10) release of the
draw cables; (11) removal of the thrust jacks and thrust beam; and (12) inspection and
repair of the outer surface of the segment.

164

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


Prestressed concrete bridges

Some of these 63 tasks are performed sequentially, while others may be carried out in parallel.
Resources are assigned to each task in terms of labour, equipment and supervision, and process
logic is assigned to each task in terms of the sequence of execution and interference with other
tasks. Instructions are formulated to model the direction of the resource ow between the
preceding and succeeding tasks. Flow-control statements are used to control task initiation, the
drawdown and release of resources, and conditional logical aspects that rule the process.

Simulation of the deck construction shows the bottlenecks in the construction process (Marzouk
et al., 2011). Cage fabrication and the availability of personnel for launch operations are two
typical bottlenecks. Sensitivity analyses can be performed to evaluate the process performance
with different combinations of resources. For example, if the rebar cage is assembled by a subcon-
tractor, the ironworkers will be available during launching, so they may be included in the
resources assigned to the launch tasks, and the simulation re-run. Process optimisation may also
include doubling the prestressing crews, or adding a grouting crew.

The alternative casting processes for the deck segments are also analysed and optimised. The
results are compared in terms of labour demand, investment in special equipment, nancial
exposure and project duration to identify the optimum casting procedure for the specic bridge
project.

3.2.5 Segment extraction


Although the cylindrical geometry of a launched bridge may resemble continuous construction
processes such as extrusion or slip forming, launching is staged with the construction of new
segments, and segment casting is followed by curing within the casting cell. At form stripping,
therefore, the new segment must be supported during its extraction from the casting cell. There
are two main problems to solve:

g to reduce the extraction friction to prevent the deck from dragging the casting cell during
launch
g to reduce the longitudinal and transverse exural stresses in the young concrete.

The outer form is stripped from the segment surface prior to launching to reduce the extraction
friction and to avoid damaging the form during launch. This can be accomplished by rotating the
outer form of the webs and side wings and lowering the bottom form table between the extraction
rails, or by lifting the segment. Lifting permits insertion of low-friction launch pads between the
segment and the bottom form table, and is the typical solution for the launch of simple spans. The
launch bearings can be mobile or xed during segment extraction, depending on whether they are
integral to the deck or to the casting cell.

g In a casting cell with mobile bearings, four hydraulic jacks at the corners of the bottom
form table are used to lift the segment. The front jacks are equipped with launch bearings
and stay in place during segment extraction. The rear jacks are used only to insert two
Neoon pads that support the rear end of the segment during extraction and slide along the
form table.
g In a casting cell with xed bearings, the launch bearings are located at the front end and at
the centre of the casting cell. Left-in-place reinforced-concrete panels protect the launch
bearings during segment casting (Figure 3.22). The segment is lifted prior to launching to
insert Neoon pads over the bearings. The rear deck overhang at the beginning of launch

165

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


Bridge Launching

Figure 3.22 Segment extraction on fixed bearings. (Reproduced with permission from ASCE)

Precast concrete panel Bottom formwork

Launch bearing Hydraulic jack

Deck

Foundation of the casting cell

and when the segment leaves the rear support is checked for the exural capacity related to
partial curing and prestressing; multiple pairs of rear bearings may be necessary to shorten
the rear overhang in long segments. Fixed bearings are frequently used in the aerial casting
cells as they localise the support reactions onto the temporary piers (Figure 3.13).

In most cases, a continuous low-friction support is used for the segment to avoid lifting. The
bottom form table includes two full-length extraction rails under the deck webs, and a central low-
ering form bridging the extraction rails. This scheme is adopted also for the mobile casting cells for
long-line casting of short segments (Figure 3.23).

Light deck segments are supported on the extraction rails by means of steel plates lubricated with
grease, hydraulic oil, soap or other such antifriction material. The high frictional resistance of
these sliding systems is compensated for by the light weight of the segment. For heavier segments,
the top anges of the extraction rails are nished with a full-length polished stainless-steel sheet,
and 1015 mm steel plates equipped with a bottom PTFE surface are aligned over them for the
entire segment length to obtain a continuous steelPTFE contact. Bakelite is a suitable alternative
to PTFE and has higher mechanical strength. In both cases thorough greasing of the sliding
surface is necessary both to reduce friction and to avoid grout penetration during casting of the
segment. The segment is match-cast onto the launch plates.

Lowering the central form table and stripping the outer forms leaves the segment on a low-friction
continuous support. The required concrete strength at form stripping is determined based on the
transverse bending in the side wings and the local stresses at the anchorages of the launch tendons
and at the load transfer points of the launch systems. Longitudinal self-weight bending rarely
governs the curing requirements on the short spans of the curing area. Deck launching extracts

166

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


Prestressed concrete bridges

Figure 3.23 Segment extraction on 50 m continuous rails

the segment from the casting cell and moves it on the rear supports of the curing area. The launch
plates of the extraction rails advance with the segment and fall on the ground in front of the casting
cell, to be cleaned and repositioned for casting of the next segment.

The vertical alignment of the extraction rails is set with extreme accuracy and checked frequently,
as the deck surface that comes in contact with the launch bearings is match-cast on top of the
extraction rails. Differences between the vertical alignment of the extraction rails and the theoreti-
cal launch surface generate hyperstatic stresses in the deck during launch and premature wear of
the Neoon pads.

The extraction rails should not deect during casting of the segment. The rails are provided
with geometry-adjustment devices to set the vertical alignment to 12 mm tolerance from the
theoretical prole. The allowed vertical tolerance may be specied as 1/5000 of the segment length
in rectilinear bridges and 1/2500 in curved bridges (AFGC, 1999). Continuous foundations are
used instead of spread footings on settling soils, and foundation piles, micropiles or jet grouting
may be combined with a light prestressing of the foundation beams for better geometry control.

Extreme precision is also required in setting the plan alignment of the extraction rails for curved
bridges, as cumulative errors in the plan radius or angle breaks at the construction joints would
result in permanent lateral offset of the deck at the piers.

Casting cells equipped with continuous extraction rails are more complex and expensive than
those equipped with xed or mobile extraction bearings. The advantages include precise vertical
alignment of the deck launch surfaces, lower stresses in the young concrete because of the continu-
ous support, and lower labour demand because segment extraction is typically unattended.

167

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


Bridge Launching

3.2.6 Curing area


The segments receive only a portion of the launch prestressing before being extracted from the
casting cell. Therefore, the casting cell is located far behind the abutment to let the segments cure
sufciently to resist full launch prestressing and the exural launch stresses on the end span of
the bridge. When possible, the curing area should be as long as the typical span of the bridge
to control backward propagation of longitudinal bending behind the abutment. The length of the
curing area is chosen also in relation to the length of the front deck cantilever during the launch
stoppages for casting of new segments, to control the irreversible deck deformations due to creep
of concrete. Control of creep deections is particularly important when the deck has spans of
constant length and the launch steps are as long as the span.

In the area between the casting cell and the launch abutment, the deck is supported on closely
spaced curing supports that are distributed according to the level of curing and prestressing
reached by the segments. The curing supports consist of reinforced-concrete pillars supporting
small launch bearings. The number of curing supports is kept as small as possible because of the
additional labour they require for the insertion of the Neoon pads during launching.

At the end of segment extraction from the casting cell, the rear end of the deck (i.e. the con-
struction joint against which the next segment will be match-cast) must be aligned with the local
radius of vertical curvature of the bridge. Angle breaks would alter the cylindrical geometry of the
deck and would cause hyperstatic stresses during launch. The distribution of the curing supports
must therefore limit the exural rotation of the rear deck section to w (1520) 103 rad.
Longitudinal bending in the curing area is the larger the longer the end span of the bridge is. If
the casting cell must be kept close to the launch abutment due to surface restrictions, one or two
temporary piers may be placed in the end span of the bridge to control backward propagation of
self-weight bending.

The rst curing support ahead of the casting cell is subject to additional restraints.

g The rear end of the deck must remain rmly supported on the extraction rails of the casting
cell, as uplift would create vertical discontinuities in the deck soft at every joint, which
would complicate launching. This requirement also affects the distribution of the curing
supports.
g The rst curing support ahead of the casting cell is equipped with lateral guides to keep the
rear end of the deck aligned with the casting cell after segment extraction. This suggests
positioning this support as close as possible to the casting cell in order to use its
foundations to resist the lateral guide forces.
g An excessive distance of the rst curing support from the casting cell complicates the
extraction of the last segment of the deck, because the support reaction provided by the
extraction rails may break the deck soft at the rear end of the segment. Damage may be
avoided by rounding off the front end of the extraction rails to progressively relieve the
support reaction, or by supporting the deck with sliding jacks in the last phases of
extraction. Sliding jacks are also used when the deck leaves the curing supports at the end
of construction (Figure 3.24).

The distance between the front-most curing support and the abutment controls backward propa-
gation of self-weight bending from the end span of the bridge, and ensures positive reactions at
the curing supports to avoid deck uplift and unexpected structural congurations. The front-most

168

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


Prestressed concrete bridges

Figure 3.24 Localised support reactions can damage the deck soffit. (Reproduced with permission from
ASCE)

curing support must be relatively far from the abutment to avoid deck uplift and to increase the
tributary deck length on the abutment when Eberspacher launchers are used.

Once the two end supports of the curing area have been positioned, curing of concrete and the
application sequence of launch prestressing govern the spacing of any intermediate supports.
The use of a large number of curing supports is discouraged by the labour demand during launch-
ing and the unpredictable distribution of support reactions due to geometry irregularities in the
deck and the supports on so short spans. When the number of curing supports is excessive, the
application of launch prestressing may be accelerated by using shorter and less overlapped ten-
dons, or a temporary pier may be inserted in the end span of the bridge to move the casting cell
closer to the abutment.

The curing supports may turn out to be excessively spaced in the nal phases of launching, when
the rear end of the deck cantilevers out by the distance between the supports. Special integrative
supports may be used only for the nal launch phases to save labour and to simplify launching.
The integrative supports are kept disengaged during standard launches by not inserting the
Neoon pads. Most curing supports require sliding jacks for the nal release of the support
reaction at the end of deck construction.

3.3. Precast segmental highway overpasses


Incremental launching of a cast-in-place PC deck requires the use of specialised equipment and the
construction of a casting yard. The cost of the casting yard depends on segment length, the cost of the
launch nose depends on deck weight and span length, and most of the investment therefore does not
depend on the length of the bridge. The savings that result from incremental launching are mostly

169

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


Bridge Launching

associated with the elimination of falsework and heavy cranes, and therefore increase with the length
of the bridge. The break-even point is reached for bridges of a specic minimum length, and incre-
mental launching tends therefore, to be more cost-effective for long bridges (Ontario, 2006).

The break-even deck length reduces signicantly when the requirements of minimum-impact
construction over highways and railways are taken into consideration. When the impact on the
travelling public is part of the scoring scheme for alternative proposals, launching may become the
rst-choice technology for new crossings over existing highways and railways, as well as for their
replacement.

The technical feasibility of launching short bridges is not in question. The challenge is to develop
short-bridge solutions that can be implemented at a competitive cost (Ontario, 2006). The exten-
sive use of launch technology for short two- or three-span crossings requires a commitment to
build a sufcient number of crossings by launching, and a modular standardised design of bridges
and construction equipment. The following guidelines may be considered:

g Design a modular system of rectilinear PC single-cell precast segmental box girders. Use
match-cast segments and epoxy joints to minimise labour demand and on-site processing of
uid concrete. Consider long-line casting of segments to provide accurate geometry and to
minimise error propagation. If the deck is skewed, keep the joints orthogonal to deck axis,
align the abutment bearings to the orthogonal, and cast skewed pier diaphragms in-place
on launch completion. If the deck must necessarily be curved in plan, rotate the cross-
section rigidly to the cross-fall, set the launch bearings along the transverse gradient of the
launch cone, and design the launch guides for the radial component of the vertical support
reaction due to the soft cross-fall.
g Use high-performance concrete to minimise the deck weight. Use powerful full-length
external tendons anchored at the end diaphragms for launch and service prestressing. Use
coupled internal gluing bars for the epoxy joints as a part of launch prestressing.
g Complete the deck behind the abutment and launch it into position in one uphill operation
with draw cables and rear thrust beams to minimise the abutment dimensions. Consider
reusable post-tensioned precast segmental foundation beams to support the launch rails in
the assembly yard. Design the launch nose and thrust systems for modularity and ease of
reuse. Minimise the cost of components that cannot be reused.
g Combine launching with lateral skidding to build the deck with less impact on the approach
embankments. To demolish the existing deck, skid it laterally over temporary pier
extensions, jack the deck to create the clearance for launching of an underbridge, and roll
the deck backward over the underbridge for progressive demolition behind the abutment.

Monolithic launching of precast segmental decks can be a very competitive solution for the
replacement or duplication of crossings over active highways and railways. Bridges with regular
standardised geometry and of a length that does not justify the investment for a self-launching
gantry are optimum candidates (Rosignoli, 2013). In large-scale projects, a precasting facility can
simultaneously feed self-launching gantry erection lines and multiple launch facilities for short
overpasses, for enhanced depreciation of investment and planning exibility.

3.4. Launch bearings and lateral guides


The technological evolution of the incremental launching construction method has seen a pro-
gressive simplication of the launch bearings for PC bridges (Rosignoli, 2000). In the very rst

170

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


Prestressed concrete bridges

Figure 3.25 Launch bearing. (Reproduced with permission from ASCE)

Neoflon pad Screw tensioning system

High-strength grout Rectangular guide frame

2 mm polished stainless-steel sheet


Concrete or steel launch plate
2030 mm elastomer block
Precast RC base plate
Lifting jack seat

applications, the deck was supported on permanent bearings that slid on stainless-steel plates
integral to the pier caps, and the length of each launch cycle was limited by the pier-cap width.
When the bearings reached the front end of the sliding plates, the launch was interrupted to lift
the deck and reposition the bearings at the rear end of the sliding plates. The cost and duration
of these operations soon suggested the use of temporary launch bearings based on Neoon pads
inserted between the deck and the xed launch bearings (Figure 3.25).

The possibility of inserting Neoon pads without lifting the deck makes the launch continuous,
much faster and less labour intensive. Load eccentricity on the piers due to the longitudinal dis-
placement of the launch bearings is also avoided. One worker is required at each launch bearing to
insert the Neoon pads and to recover and clean the pads expelled frontally from the bearing.
Automatic insertion systems have been tested for support reactions up to 2.0 MN (AFGC,
1999) to reduce the labour demand of launch operations.

The Neoon pads are made up of elastomer sheets reinforced with steel shims and glued to a lower
PTFE plate. The elastomer spreads the vertical load and absorbs surface irregularities, and its
high friction coefcient against the concrete surface of the deck drags the pads along the launch
bearings during launching. The bottom PTFE plate reduces the sliding friction, and may be
dimpled for additional lubrication with silicon grease. The steel shims stabilise pad geometry
under the longitudinal shear. The design of the pads depends on the elastomer hardness, plan
geometry, shape factor, thickness of elastomer layers, number and distribution of steel shims,
compressive stress level, and required rotational stiffness. Neoon pads made with elastomers

171

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


Bridge Launching

Figure 3.26 Launch bearing assembly. (Photo: URS)

having lower hardness ratings offer some advantages over those made from harder elastomers, and
lower rotational stiffness.

Over the years a large number of different details for the launch bearing assembly have been tested,
not always with satisfactory results. The recommended assembly consists of four elements
(Figure 3.26). From bottom to top, these are:

g A precast reinforced-concrete base plate adhered to the pier cap with high-strength grout.
Levelling screws at the four corners facilitate alignment with the local launch gradient. A
steel frame may be used as a left-in-place form for the perimeter of the base plate, to carry
the levelling screws and for concrete connement. No anchor dowels are used between the
base plate and the pier cap, as the friction coefcient between two concrete surfaces is one
order of magnitude higher than that of the steelPTFE interface.
g A 2030 mm elastomer block, laminated or full-depth, supported on the base plate to
improve the stress distribution and rotational capacity of the assembly.
g A 50 mm steel plate rounded at the transverse edges to facilitate longitudinal insertion and
expulsion of the Neoon pads. The plate is welded to a guide frame that hugs the
elastomeric block and base plate to prevent lateral displacements of the launch plate. The
top surface of the plate is mechanically attened to ensure planarity tolerances of a few
tenths of a millimetre.
g A 2 mm stainless-steel sheet, mechanically polished to 2 mm roughness, which is bolted to
the steel plate on the rear side (the insertion side of the Neoon pads) and tensioned with
screw systems on the expulsion side of the Neoon pads. The launch bearings allow free
deck displacements in the transverse direction and do not provide lateral guidance.

In some applications, the launch plates have been assembled onto permanent bearings and
removed on launch completion (Figure 3.27) (Cremer et al., 2003; Llombart and Revoltos,

172

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


Prestressed concrete bridges

Figure 3.27 Launch plate applied to a permanent bearing. (Photo: URS)

2000). When permanent elastomeric bearings are used to support launch plates and lateral launch
guides, the lateral exibility of the bearings may be used to reduce the transverse loads applied to
the piers (Bennett and Taylor, 2002). Prior load testing of elastomeric bearings for shear stiffness
allows evaluation of the horizontal loads applied to the piers by measurement of the shear
deformations of the bearings. The launch plates may also be assembled on hydraulic jacks fed
by a closed-loop system designed to create a plastic torsional hinge that controls or minimises the
transverse torque applied to the pier (Figure 3.28). This solution is sometimes adopted on steel
temporary piers to equalise the load in the columns.

Decks with varying width or plan radius may require lateral skidding of the launch bearings
during the course of launch. Skidding bearings may be lodged into sledges that also contain the
lifting jacks for maintenance of the launch surfaces. Full-width launch sledges are used when
incremental launching is combined with lateral skidding (Figure 3.11). When the deck is curved
in plan and the launch nose is rectilinear, wide launch bearings are used to accommodate lateral
nose displacements during lunching. When tight plan curves bring the nose too close to the edges
of the pier cap, the nose is made narrower than the deck, and special launch bearings are used for
the nose in the central area of the pier cap (Figure 3.29). The nose bearings are no longer necessary
after arrival of the PC deck, and may be repositioned from pier to pier during the stoppages in the
launch to cast the segments.

The contact surface of the launch bearings is designed for the peak SLS support reaction Rmax
determined for self-weight only. The effects of irreversible creep deections of the deck, geometry
imperfections, misaligned launch bearings and thermal gradients are rarely considered in the
analysis of Rmax because of the large margins of safety of the elastomer from rupture. The design
pressure is kept as high as possible to reduce the friction coefcient of the steelPTFE sliding
surface and to use small Neoon pads for easier handling. Setting bnf as the transverse width and
Lnf as the length of each pad, the SLS peak vertical compressive stress snf in the elastomer is

173

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


Bridge Launching

Figure 3.28 Hydraulic launch bearings. (Reproduced with permission from ASCE)

Figure 3.29 Nose bearings for curved bridges. (Photo: Torroja)

174

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


Prestressed concrete bridges

dened as

Rmax
snf = 1315 MPa (3.11)
bnf SLnf

The Neoon pads are generally 300500 mm wide, and may be designed to be 1020 mm narrower
than the extraction rails of the casting cell in order to work within the prints of the extraction
plates, where the deck surface is particularly regular. This solution is particularly advantageous
when the rest of the deck surface has board-marked nishing. Generous 458 chamfers are provided
at the soft edges of box girders and ribbed slabs to keep the support reaction aligned with the web
axis. The edge of the Neoon pads should not be closer than 100 mm to the deck edge in order to
avoid damage to the unreinforced deck corner.

The exural rotations of the deck cause vertical compressive stress gradients in the launch bear-
ings. The stress gradients depend on the length of the bearings in a quadratic ratio, and SLnf
should therefore be kept as short as possible. Long launch bearings are supported on thick
multi-layered elastomeric bearings that improve the rotation capacity of the assembly, reduce the
stress gradients and avoid excessive wear of the Neoon pads.

The length Lnf of the Neoon pads is chosen according to preference. Short pads are light and easy
to handle, but require frequent insertions and higher attention from the workers. Longer pads
facilitate positioning (especially when defects in the deck surface tend to rotate them) and the
introduction of shim plates between the pads and the deck, but are heavier and may require two
workers for handling.

In the transverse direction, the launch bearings are 4060 mm wider than the Neoon pads so as
not to damage the pads in the case of eccentric insertion. In the longitudinal direction, the bearings
have 100150 mm rounded ends that facilitate the insertion and expulsion of the Neoon pads and
minimise stress concentrations. The total length of the launch bearings is, therefore,

Lb = 0.30 + SLnf = 1.21.5 m (3.12)

The kinetic friction coefcient of new dimpled PTFE plates over polished stainless steel is typically
in the range 14% in relation to lubrication, compressive stress and temperature. The breakaway
friction is higher after a weekly or bi-weekly launch stoppage because of the tendency of PTFE to
seize to the stainless-steel surface.

Lateral guides are used alongside the launch bearings to keep the deck guided during launch and
to transfer lateral forces. Even in the absence of macroscopic geometry imperfections or cross-fall
of the deck soft, the deck tends to drift to one side during launching. In non-seismic areas,
rectilinear decks are guided at the lead pier and at the front end of the casting cell. Multiple guides
are used in seismic areas and for curved bridges to minimise lateral bending in the deck. The lateral
load depends on the vector decomposition of the thrust force and the transverse exibility of the
piers, and rarely exceeds 10% of the vertical support reaction.

The lateral guides produce xed or adjustable guide forces through steelPTFE contacts. The
guides are kept in position with steel brackets anchored to the pier walls or to the pedestals of
the launch bearings with through prestressing bars. When the launch bearings are equipped with
skidding shoes for lateral deck displacement on launch completion, the launch guides are

175

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


Bridge Launching

Figure 3.30 Adjustable launch guide. (Reproduced with permission from ASCE)

Jack

Pier
Jack support bracket

incorporated in the framing system of the support saddles to restrain the deck also during skidding
(Figure 3.23). Hydraulic guides are more expensive but allow lateral deck displacement during
launching (Figure 3.30). The lateral guides are almost always located at the bottom slab level for
optimal load dispersal within the deck. Special offsets are used in the rounded cross-sections
(Rosignoli, 1998a).

3.4.1 Pier-cap gear


The pier caps of the permanent and temporary piers for launched bridges are designed for the
technological requirements of launching, and to provide adequate working areas and easy access
to launch bearings and lateral guides. The typical pier-cap gear includes two launch bearings, two
lateral guides, Neoon pads for bearings and guides, and lifting jacks for maintenance of the
launch bearings and replacement of defective Neoon pads.

The pedestals of the launch bearings are designed for the depth required for easy operations and
the deck cross-fall. Clearance is kept behind and ahead of the pedestals for operation of the lifting
jacks, which are also located under the deck webs. The lifting jacks are kept 70100 mm from the
edges of the pier cap to avoid damaging the unreinforced concrete cover in the corner. Wider clear-
ance may be necessary when the lifting jacks are placed on PTFE plates to recover pier deections.
Long Neoon pads may cause difculties at deck lifting, as they remain trapped between the jacks
and the deck and must be cut.

Jacking may be necessary during the launch, and is always necessary on launch completion to
replace the launch bearings with the permanent bearings. Jacking for launch bearing maintenance

176

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


Prestressed concrete bridges

is limited to a few millimetres, and light aluminium at jacks may be used for easier handling under
the deck. Jacking may be longer on launch completion, especially at the end abutment and the last
pier, where it is necessary to compensate for creep deections accumulated in the front deck region
during launching. The support reactions are weighted and adjusted at every pier during deck
placement on the permanent bearings to minimise stress redistribution due to time-dependent
effects. At the temporary piers it is also necessary to relieve the support reaction at the end of
launching, and longer-stroke jacks are used with hydraulic systems designed to provide volumetric
control on synchronised displacements.

The jacks have tilt heads that allow the distribution plates to adapt to the deck surface, which is
almost always inclined. The use of wedge shims is not recommended because of the risk of plate
expulsion. The jacks are provided with mechanical ring nuts that limit the hydraulic support to
the phases of lifting and lowering. The jacks may be placed on low-friction surfaces (polished
stainless-steel plates on PTFE sheets) to avoid shear transfer due to residual pier deections caused
by launch friction and the launch gradient. As the steelPTFE support of the jacks is cleaner and
less worn than the Neoon pads used for launching, the pier recovers part of the longitudinal
deection at lifting. These movements must be accounted for when setting up the batteries of
lifting jacks, to avoid cracking of the pier-cap edge.

The use of lifting batteries composed of many small jacks instead of a few stronger units facilitates
jack handling under the deck and results in less restrained lifting schemes. When the pier cap is too
small to align the jacks under the webs, steel brackets may be anchored to the pier walls by means
of dry joints with distributed shear keys and through prestressing bars to support the jacks.

Launch bearings and lateral guides are located at the edges of the pier cap, and the central area is
available for the hydraulic systems and power pack unit of the lifting jacks and to store jacks,
grease, Neoon pads and the bridge bearings. The permanent bearings are lifted onto the pier cap
prior to deck arrival, using ground cranes or a service hoist applied to the launch nose (Rosignoli,
2013). This solution is particularly effective for tall piers and heavy and voluminous bearings of
railway bridges. When the launch nose is narrower than the deck soft, four launch bearings are
required at each pier (Figure 3.29). When the piers are very slender longitudinally, small
reinforced-concrete corbels may be cast onto the pier caps to recover the residual pier deection
on launch completion by jacking against niches left in the deck soft ahead of the jacking point.

3.4.2 Effects of vertical misalignment of launch bearings


Vertical misalignment of supports generates hyperstatic stresses in a continuous beam. Ribbed
slabs with a double-T section and solid and voided slabs are relatively exible, while the higher
exural stiffness of PC box girders makes them more susceptible to the effects of bearing
misalignment.

In the long-term, a +10 mm misalignment between the supports of a 4050 m span may be tolerated,
as creep of concrete reduces and redistributes the hyperstatic stresses. During launching, the stresses
due to bearing misalignment depend on the instantaneous elastic modulus, and may cause cracking
and deck uplift. Several factors may generate hyperstatic stresses in relation to poor geometry control:

g Errors in the elevation of the launch bearings due to inaccurate initial positioning or to
settlement of foundations or temporary piers under the load applied by the deck. These
errors may be easily corrected by shimming.

177

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


Bridge Launching

g Geometry imperfections in the deck. Imperfections may result from inaccurate vertical
alignment of the extraction rails of the casting cell. When these errors are discovered, the
extraction rails are realigned but the geometry irregularities in the portion of the deck
already built cannot be corrected. Because of the repetitiveness of segmental construction,
these errors affect all deck segments at the same location. Geometry imperfections may also
result from angle breaks at the construction joints between match-cast segments.
g Unforeseen events, such as breaking of a launch bearing or the expulsion of a Neoon pad.
Such events can be avoided by supervising the launch and training the operating crews.

The effects of the vertical misalignment of launch bearings are different in the longitudinal and the
transverse directions. In the longitudinal direction, the launch prestressing is designed to allow
deck lifting at every support to replace the Neoon pads if required. A +10 mm displacement
is typically applied to every launch bearing, with the other bearings assumed aligned, and the
stresses are enveloped. The 10 mm displacement derives from the assumption of 7 mm misplace-
ment and 7 mm settlement, combined with the square root of the sum of the squares rule, and may
therefore be reduced in the case of stiff foundations and simple deck geometry, or increased in the
case of a curved deck and exible foundations. A +10 mm uplift is considered at a launch abut-
ment equipped with Eberspacher launchers, and a +5 mm displacement is considered at the
leading end of the casting cell, because of the exibility of the extraction rails.

Launch prestressing is often designed for the lower-edge tensile stress at midspan, which is not
greatly affected by the misalignment of supports. The variation in the negative moment in a
support section subject to a vertical displacement Wm is
EI
DM = k Wm (3.13)
L2
with 3 , k 6 in relation to the location of the cross-section within the continuous beam. As the
span length reduces negative bending in a quadratic ratio, the axial stress due to launch-bearing
misalignment is often tolerable without any increase in the launch prestressing.

In the transverse direction, launch bearings at the wrong elevation cause transverse eccentricity in
the support reaction. An eccentric load applied to an elastic beam produces torsion and distortion,
and the high torsional stiffness of a PC box girder may lead to bearing decompression/uplift and
doubled support reaction on the adjacent bearing. An eccentric load also causes transverse bend-
ing in piers and foundations. The lateral exibility of the piers reduces torsion and distortion in the
deck, and may be considered in the analysis.

During the service life of a bridge, the torsional stiffness of the box girder controls the tangential stress
due to eccentric live loads, and the frame stiffness of the cross-section along with the in-plane exural
stiffness of slabs and webs rapidly reduce distortion (Rosignoli, 1981). In most cases, therefore,
important torsional effects can arise only in the presence of eccentric localised loads. Although the
highest localised loads applied to a bridge deck are the support reactions, their transverse eccentricity
is generally disregarded in the design. Bearing failure is improbable, in-place casting of the deck or the
bearing seats avoids transverse support eccentricity, and the pier diaphragms prevent distortion.

In a launched bridge, the deck sections over the piers are devoid of diaphragms, and misaligned
launch bearings may cause distortion and severe overloading. In the support section shown in
Figure 3.31, the distances W3 and W4 of the support points from the theoretical launch surface

178

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


Prestressed concrete bridges

Figure 3.31 Load imbalance due to misaligned launch bearings. (Reproduced with permission from
ASCE)

1 2

4 3
W4 W3
Theoretical launch surface

b
R4 R3

dene the vertical misalignment of the launch bearings. Misalignment may be divided into a sym-
metrical error Wm , the effects of which are only longitudinal, as per Equation 3.13,

W4 + W3
Wm = (3.14)
2

and in a hemi-symmetrical error Wh

W4 W3
Wh = (3.15)
2

which causes a rotation wx = 2Wh /b of the bottom slab in the transverse plane. The effects of Wh
are more pronounced than the effects of Wm , as the cross-section is not prestressed in the trans-
verse direction and the distance between the launch bearings is much smaller (i.e. rotations are
amplied). Large rotations and high torsional stiffness cause large differences between the support
reactions of the adjacent bearings (Rosignoli, 1997b), the overloaded bearing may break
(Figure 3.32) and the cross-section distorts.

In a rectilinear single-cell box girder, the effects of Wh deviate from Bredts theory. None of the
assumptions in this theory (rigid in-plane rotation, torsion applied through the theoretical distri-
bution of tangential stress, and free warping) is respected, and the deformation is a combination of
rigid rotation and distortion (Rosignoli, 1998d). Rigid rotation depends on the torsional stiffness
of the cross-section, which limits its capacity to rotate about the longitudinal axis. Distortion
depends on the interaction between the frame stiffness of the cross-section and the in-plane stiff-
ness of the web and slab plates.

If the x-axis corresponds to the deck axis, ht(x) denes the vertical displacement of the deck lower
nodes due to torsion, and hd(x) denes the displacement due to distortion. Both components are
hemi-symmetrical (i.e. they are equal and opposite in sign at the two nodes 3 and 4 in Figure 3.31).
Section O over the misaligned launch bearings is taken as the origin of the frame of reference, and
the other support sections are assumed to be aligned. In order to maintain contact with the launch
bearings,

hd(0) + ht(0) = Wh,O (3.16)

179

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


Bridge Launching

Figure 3.32 Collapsed launch bearing

The distortion hd of the cross-section is dened by a differential equation similar to the one for a
beam on elastic foundation (BEF) (Wright et al., 1968):

d 4 hd
ECd + khd = ph (3.17)
dx4

where Cd is a cross-sectional property related to warping, k is the deformational stiffness of a unit


length of the box cell, and ph is a hemi-symmetrical distributed load. The property of the cross-
section that corresponds to the BEF moment of inertia is dened by evaluating the share of
hemi-symmetrical load resisted by the webs and slabs through in-plane bending, and can be
expressed as

b  
Cd = k8 k2 k7 + k3 k8 + 2k4 k9 (3.18)
4

with the following coefcients

b2 tb + c(a + 2b)tw
k1 =   (3.19)
(a + 2d )3 /a tt + c(2a + b)tw
 
(a + 2d )3 b c 
k2 = a k1 tt tb 1 k1 tw (3.20)
6a2 4 2

b2
k3 = t (3.21)
12 b

180

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


Prestressed concrete bridges

b c 
k4 = c tb + 2 k1 tw (3.22)
4 6
 
(a + 2d )3 b c a
k5 = a tt + tb + 1 + tw (3.23)
6ab 4 2 b

b c a
k6 = c tb + 2 + tw (3.24)
4 6 b

k6 k5

1 c a
k7 = 2h + (3.25)
a ab
k5 k3 k6
c
k6 k5

k8 = 2h c a (3.26)
ab
k5 k3 k6
c

k6 k5
h
c a
k9 = 1 + (a b) (3.27)
c ab
k5 k3 k6
c

The vertical shear force V produced in the middle of the bottom slab by the frame action of the
cross-section under a hemi-symmetrical load ph = 1 is

a4 b2 c
+ 3 (2a + b)
t3t tw
V=  3  3  (3.28)
a b 2c  2 
(a + b) + + 3 a + ab + b 2
tt tb tw

and the cross-sectional property that corresponds to the BEF modulus of foundation is

2(a + b)E
k=      (3.29)
  a 4
a cb2 2b2
1 y2 V + 3 (2a + b)V
t3t a+b tw a + b

The BEF is solved as a beam of innite length. Aligned launch bearings at the adjacent piers do
not provide support to the BEF, as in the absence of pier diaphragms the cross-section can distort.
In the origin section O, vertical misalignment of the launch bearings unbalances the two support
reactions (Figure 3.31). These may be divided into a symmetrical component Rs,O = (R4 + R3 )/2,
which generates longitudinal bending and shear, and a hemi-symmetrical component Rh,O =
(R4 R3 )/2, which is applied to the BEF to analyse distortion, and which generates a torsional
moment TO = Rh,Ob.

If the deck cross-sections were rigid, the torsion diagram due to the moment TO would be limited
to the two spans supported at section O. If the two spans had the same length, two equal moments
at the adjacent support sections A (the bearings of which are assumed to be aligned) would
balance the torque at section O, so that TA = TO /2. However, TA is also produced by

181

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


Bridge Launching

Figure 3.33 Loading of the BEF. (Reproduced with permission from ASCE)

Rh
B A O A B
x

RhB RhA z RhA RhB


L L L L

hemi-symmetrical components of the support reaction, TA = Rh,Ab, and the support sections A
also distort. Rigid rotations occur at sections A to compensate for distortion and maintain contact
with the launch bearings, and a portion of torsion migrates toward the external spans. Torsion
migration theoretically affects the whole deck; in practice, high values of Cd and k fade the distor-
tion rapidly, and torsion spreads only to the rst two spans on either side of section O. Therefore,
the BEF is loaded as shown in Figure 3.33.

The characteristic length of the BEF (Hetenyi, 1971) is



4 k
l= (3.30)
4ECd

The equation for the right branch (x 0) of the BEF elastic line produced by a load Rh,O applied
to the origin is

Rh,O l lx
hd (x) = e (cos lx + sin lx) (3.31)
2k

The deection is generally negligible beyond the rst zero point at x = 3p/4l. In a PC box girder,
distortion fades within about half a span, and the effects of the hemi-symmetrical components
of the support reactions do not superimpose. At sections A, distortion is caused only by the
hemi-symmetrical component Rh,A of its own support reaction. Therefore, from Equation 3.31
one obtains the hemi-symmetrical deection hd,A = Rh,Al/2k and the rotation of the bottom slab
wx,A = Rh,Al/bk.

As the launch bearings are assumed to be aligned correctly at section A, a rigid torsional rotation
qA = wx,A must occur at section A to maintain contact with the launch bearings. The difference
between the rigid rotations qA and qB of the support sections A and B is equal to the torsional
rotation of the span AB under the torsional moment TB = Rh,Bb. Imposing this condition, one
obtains

l b
+ L
bk Ct
Rh,A = Rh,O (3.32)
4l 2b
+ L
bk Ct
l
Rh,B = bk Rh,O (3.33)
4l 2b
+ L
bk Ct

182

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


Prestressed concrete bridges

where L is the length of the span, and the torsional stiffness Ct , neglecting restrained warping, is
dened by the second Bredt equation (m = 1/n, where n is Poissons ratio):
m
E [(a + b)h]2
2(m + 1)
Ct = (3.34)
a b c
+ +2
tt tb tw

The equation for distortion thus becomes (x 0)

l
hd (x) = [R elx (cos lx + sin lx) Rh,A el|Lx| (cos l|L x| + sin l|L x|)
2k h,O
Rh,B el|2Lx| (cos l|2L x| + sin l|2L x|) + Rh,B ] (3.35)

Finally, the displacement ht(x) due to torsional rotations is (in the two spans)

b2 1
h0A
t (x) = (L x)Rh,O + Rh,B L (3.36)
2Ct 2

b2
hAB
t (x) = (2L x)Rh,B (3.37)
2Ct

With dened hd(0) and ht(0) for the hemi-symmetrical force Rh,O , the calculation of the actual
hemi-symmetrical component of the support reaction based on the assumed launch bearing
misalignment is immediate. In the absence of a bidirectional restraint, this force cannot exceed the
reaction Rs,O produced by each bearing in balanced conditions. When 50% of the continuous
beam support reaction is reached, one bearing carries the whole reaction and the deck lifts from
the adjacent bearing, avoiding further deformations.

Setting

4 3k
ls = 3 (3.38)
Ec tw C9

the longitudinal axial stresses generated by distortion at the web corners are

3Rh,O
s3 =   (3.39)
1 + k1 ls tw c2

and

s2 = k1s3 (3.40)

These stresses superimpose on the longitudinal axial stress generated by restrained warping.
Finally, transverse bending in the cross-section, dened by Equation 3.28 for a couple of unit
forces, is obtained by linear proportion with the actual hemi-symmetrical component.

The above approach has been validated using nite-element analysis (Rosignoli, 1998b). With a
web inclination up to 208, the average error is 2.3%. Finite-element analysis also conrms the
damping distance of Wd(x). A greater web inclination causes a larger error because the shear
gradients due to restrained warping reduce the distortion. Analysis has shown that only one-half

183

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


Bridge Launching

of the vertical misalignment of the launch bearings is resisted by torsion; the remainder is resisted
by distortion, and causes signicant longitudinal and transverse bending and shear.

A 10 mm vertical misalignment of one launch bearing modied the support reactions calculated for
aligned bearings with hemi-symmetrical components of about 40%. A 25 mm misalignment (which
may result, for instance, from the expulsion of a Neoon pad) may decompress the launch bearing,
thus doubling the load on the adjacent bearing. These results conrm the load differences measured
in the eld. Monitoring of the Amiens Viaduct (Godart et al., 1989) showed hemi-symmetrical com-
ponents of the support reactions of about 20%, and monitoring of the Champigny-sur-Yonne
Bridge showed that only one of the two abutment bearings was actually loaded.

3.4.3 Local stresses at the launch bearings


The bottom webslab node of a PC box girder is subject to complex stress distributions during
launching.

g Nosedeck interaction in the front deck region, continuous beam behaviour in the rear
deck region, launch prestressing, misaligned launch bearings, geometry irregularities,
thermal gradients and irreversible creep deections generate axial and tangential stresses in
the longitudinal direction.
g Thermal gradients and torsion and distortion due to misaligned launch bearings and
geometry irregularities generate axial and tangential stresses in the longitudinal and
transverse directions.
g Dispersal of the support reactions within the webs generates vertical compressive stress in
the deck support regions.
g Sliding friction and the launch gradient generate longitudinal axial stress in the deck
support regions. These stresses are localised at the bottom of the webs and jeopardise the
conservation of planar sections.

Most of these stresses may be signicant. The state of stress in the cross-section is checked using
serviceability criteria in relation to cracking of concrete and the SLS stress limits specied by the
design standards for PC sections, and using strength criteria in relation to the cross-section
capacity at the ultimate limit state (ULS).

3.4.3.1 Vertical axial stress


The support reactions transferred by the launch bearings are not much smaller than the loads
transferred by the permanent bearings on launch completion. The permanent support reactions
are applied to the deck in specic locations, and pier diaphragms are typically provided at the
support sections to prevent distortion of the cross-section and to facilitate dispersal of localised
loading. The support reactions are applied throughout the length of the deck during launching,
and the cross-sections are devoid of any stiffening. It is therefore necessary to check the stress
dispersal within the webs in the longitudinal, vertical and transverse direction.

The launch support reactions should be applied in the midplane of the webs to avoid local bending.
The exural stresses due to eccentric support reactions are the higher the more slender are the webs
and bottom slab. Webs of 0.50 m thickness will barely experience any cracking due to transverse
bending, but 0.30 m webs may require a wider bottom slab to move the Neoon pads outward and
align the support reaction. When the webs are inclined, a vertical edge may be used at the bottom of
the webs to apply the support reaction beneath the geometric webslab node (Figure 3.34).

184

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


Prestressed concrete bridges

Figure 3.34 Lower node schemes. (Reproduced with permission from ASCE)

e e

The soft zones that come into contact with the Neoon pads during launching are designed as
horizontal crosswise. Inclined sliding surfaces cause lateral drift during launch and overload the
lateral guides, and, for aesthetic reasons are only used for the launch of short-span solid and
voided slabs. For inclined sections, the soft scheme shown in Figure 3.35 is typically adopted.
These regions of the bottom slab are match-cast over the extraction rails of the casting cell, which
are also horizontal crosswise.

The outer edge of the Neoon pads is kept 70100 mm from the outer edge of the bottom slab to
avoid damaging the unreinforced concrete cover in the corner. This may require widening of the
bottom webslab node. The tendon ducts for internal parabolic prestressing are empty during
launch and reduce the vertical-load-dispersal capacity of the webs; the American Association of
State Highway and Transportation Ofcials (AASHTO, 2014) species a concrete cover of at least
150 mm to the underside (Figure 3.36).

Figure 3.35 Soffit arrangement for an inclined bottom slab

185

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


Bridge Launching

Figure 3.36 Clearance of Neoflon pads. (Reproduced with permission from ASCE)

Empty ducts

Neoflon pad

Min. 150 mm

70100 mm Min. 20 mm

Finite-element analysis conrms 458 dispersal of the support reactions within the webs. The sharp
variations in the vertical axial stress in the proximity of the launch bearings are taken into account
in the local stress verications.

3.4.3.2 Longitudinal axial stress


The load that the Neoon pads apply to the deck is not orthogonal to the support surface. In the
presence of a launch gradient, the vertical support reaction is inclined relative to the deck gravity
axis. Sliding friction further rotates the resultant and produces an additional longitudinal load
component. These forces are easy to determine starting from the local launch gradient and the
distribution of the continuous beam support reactions. The longitudinal axial stress may be high
prior to dispersal within the cross-section, and adds to the compressive axial stress generated by
general negative bending and launch prestressing.

At the lower edge of the support sections, the longitudinal axial stress deviates from the Navier
distribution and reaches higher values. Deck monitoring and nite-element analysis conrm
out-of-plane deformations of the cross-section when passing over the launch bearings. These local
stresses dissipate within a distance comparable to the deck depth. The tangential stress distribution
is also difcult to evaluate theoretically. Finite-element analysis conrms high stress gradients that
dissipate within a distance of half the deck depth.

The longitudinal axial stress is higher in the proximity of the Eberspacher launchers, as the thrust
force that these devices impart to the bottom web-slab node is much higher than the frictional
resistance of the launch bearings. The lifting jacks of the launchers are equipped with tilt heads
to improve the adhesion of the contact plates to the deck surface. The surface of the contact plates
is knurled to dissipate local stress peaks, and is as large as possible to reduce the mean contact
stress. The support jacks of the most advanced Eberspacher launchers are equipped with
proximity sensors that provide feedback to a programmable logic controller in order to reduce the
lifting speed at contact to avoid impacts.

Careful design is indispensable when Eberspacher launchers are placed on the piers. When the
launchers are at the abutment, the deck spans supported on the launchers are shorter than the full
pier spacing, and the thrust force combines with a smaller support reaction and less negative bend-
ing. The support reactions are higher on the piers, and higher launch loads combine with full-span

186

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


Prestressed concrete bridges

negative bending. Synchronisation is also important when multiple pairs of launchers are used to
move the deck, in order to minimise launch load imbalance.

3.4.3.3 Stress checks


When passing over launch bearings and Eberspacher launchers, the lower region of the webs is
subject to a triaxial state of stress. If the launch bearings are properly aligned and the temperature
gradients are low, the transverse axial stress sy due to bearing misalignment is smaller than the
longitudinal axial stress sx due to launch prestressing and negative bending, and the vertical axial
stress sz due to the dispersal of the support reactions within the webs. The Eberspacher launchers
are set to provide equal support reactions to the two webs independently of the lifting stroke,
which avoids transverse bending in the deck. In both cases, therefore, the state of stress can be
assumed as planar, and the principal stresses can be derived from Mohrs circle:
 
s1 s + sz 1  2
= x + sx sz + 4t2xz (3.41)
s2 2 2

where txz is the tangential stress in the xz plane. The principal stresses are checked against concrete
strength criteria based on biaxial compression. For instance, according to the CEB/FIP Model Code
1990 (CEB/FIP, 1993), under the assumption that the biaxial compressive strength is 20% higher
than the uniaxial one, setting a = s1 /s2 , where s1 is the largest principal stress at failure and s2
is the smallest one, the strength of concrete under biaxial compression may be estimated as

1 + 3.80a  
s2 = fck + 8 (3.42)
(1 + a)2

where fck is the 28-day characteristic compressive strength of concrete (in mega-pascal, on cylin-
ders). Under the assumptions that the SLS load lines have the same shape as the failure curve, and
that the elastic limit is 0.6fck under axial compression, it is possible to determine the biaxial elastic
domain of concrete in the s1s2 plane. Then, the SLS state of stress in the lower region of the webs
is compared graphically with the stress contours. Stresses applied many times and for long periods
of time may cause irreversible creep strains; in most cases, however, the long curing time of con-
crete limits this effect.

3.5. Time-dependent effects


In the design of reinforced-concrete structures, concrete is considered linearly elastic, although in
reality it has an inelastic and time-dependent behaviour. This model leads to simple design
methods but ignores phenomena that may cause SLSs and rapid structure deterioration, and may
therefore lead to an incorrect evaluation of the state of stress and the safety margins of a structure.

The time-dependent behaviour of concrete became evident as PC bridges built as balanced cantilevers
with midspan hinges became more widespread. This construction method requires the application
of eccentric prestressing at short curing, and midspan hinges prevent bending redistribution from
the negative bending regions to the midspan regions. This resulted in large time-dependent deec-
tions of the cantilevers and permanent angle breaks in the midspan hinges.

Although it uses the same materials, the incremental launching construction method generates
much smaller time-dependent effects. These effects may be analysed by superimposing the effects
of the shrinkage and creep of concrete on the effects of relaxation of the prestressing steel,
although these effects are related to and inuence one another.

187

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


Bridge Launching

Time-dependent deformations occur throughout the length of a PC launched bridge and are larger
in the front deck region, because of the concomitant higher launch stresses and the longer
exposure to time-dependent effects. Creep deformations have minimal inuence on the bridge
during its service, as the support reactions are weighted and corrected during deck positioning
onto the permanent bearings, correct support reactions minimise the subsequent time-dependent
stress redistribution, and integrative prestressing reduces permanent bending and shear.

During launch, irreversible time-dependent deck deections cause stress redistribution and
hyperstatic effects. Complexity of the load history and variability of the creep function of real
concretes complicate the prediction of these effects. Many structural analysis programs implement
time-dependent formulations of materials, and the problem can be studied for non-cracked sections,
as launch prestressing is typically designed for no edge decompression under self-weight. Specic
ageing viscoelastic models have also been formulated for launched bridges to follow the concrete
strain history due to the application of launch and integrative prestressing, imposed displacements,
and delayed restraints to deck segments of different age during launching (Mapelli et al., 2006).

In most practical cases, the time-dependent effects are controlled through proper design. The cast-
ing cell is positioned so as to avoid a long front cantilever during the launch stoppages for casting
of new deck segments, the launch steps are designed to avoid repetitive structural congurations
that amplify the irreversible creep deections in the front deck region, and launch prestressing is
designed by keeping a margin from the SLS tensile stress limit specied by the design standards to
account for the additional edge stresses due to the hyperstatic effects of the non-launchable deck
prole. In the most delicate cases, the support reaction applied to the launch nose is adjusted by
shimming the Neoon pads, or by placing the launch bearings on hydraulic jacks.

3.5.1 Shrinkage
Shrinkage is a slow, load-independent concrete deformation due to chemical and physical
reactions related to the cement setting. In spite of the slow development, one-third of the total
strain may occur in the rst 15 days of curing, and in the rst year it may reach 80%.

Shrinkage depends on the mechanical properties of the concrete, the quantity and quality of
cement and admixtures, the water-to-cement ratio, the aggregate-to-cement ratio, the aggregate
rigidity, and the humidity and temperature of the curing environment (Collepardi, 1980). An
increase in the water-to-cement ratio increases shrinkage because of the higher quantity of water
that can evaporate and the higher porosity of the concrete, which facilitates its evaporation. A
reduction in the aggregate-to-cement ratio increases shrinkage because of the higher volume of
cement, which is responsible for this phenomenon, and of the lower volume of aggregates, which
oppose the contraction of volume.

In a generic instant t, the shrinkage strain 1cs(t,t0 ) measured from a reference instant t0 can be
expressed as
1cs(t, t0 ) = 1cs,1k1(t, t0 ) (3.43)
where the nal value 1cs,1 and the timedevelopment function k1(t,t0 ) (which tends to a long-term
unit value, more rapidly for thinner elements) are established by the design standards. The
evolution of concrete technology is modifying the development of shrinkage over time, and the
most recent design standards specify values for 1cs,1 and k1(t,t0 ) that may differ signicantly from
the values seen some years ago.

188

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


Prestressed concrete bridges

Shrinkage is higher on surfaces exposed to the air and lower inside concrete, because of the smaller
variations in humidity and the presence of reinforcement, which opposes these deformations.
Although creep of concrete rapidly reduces the strain gradients, in the rst curing phases these can
exceed the tensile strength of concrete and generate diffused supercial cracking, which may cause
durability issues if not controlled properly. Similar issues arise at the webslab nodes of a box
girder and, if webs and slabs have different thickness, differential shrinkage may worsen supercial
cracking.

The segments of launched bridges are cast in an accessible casting cell, and most of the parameters
that govern shrinkage can be easily controlled. The water-to-cement ratio may be reduced by using
admixtures, by batching concrete in-place, and by placing it using buckets or conveyor belts. The
segment (one-phase casting) or the top slab (two-phase casting) are typically cast on a Friday and
the forms are stripped on the Monday, which ensures 2 days of protected curing without any
impact on the cycle time. The casting cell is often sheltered for additional control of temperature
and humidity (Figure 3.37).

The upper region of the top slab is a critical region in terms of durability because penetration of
chloride and other aggressive agents may cause depassivation and may trigger corrosion of the
reinforcement. Wide uncontrolled cracks in the concrete cover may also trigger freezethaw cycle
deterioration. Further local reduction in the water-to-cement ratio may be achieved by screeding
the top slab (Figure 3.37). Screeding mechanically expels the bleeding water, which increases the
compactness and tensile strength of the concrete cover and reduces its permeability, giving
enhanced protection of the top reinforcement. Screeding is easy and inexpensive in a xed casting

Figure 3.37 Rail-mounted screed in sheltered casting shell

189

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


Bridge Launching

cell, and also reduces the number of workers required for concrete nishing, which soon covers the
cost of the screed.

If necessary, draining felts may be applied to the outer form of the casting cell to reduce the
porosity of the concrete cover throughout the outer deck surface, to increase the supercial
strength, to protect the surface during the rst curing phases and to avoid the use of form oil.

3.5.2 Creep
Creep of concrete is the spontaneous growth over time of the deformation produced by a constant
load. If a concrete specimen is subject to a constant compressive stress s0 from an instant t0
measured from its casting, its shrinkage-free deformation in a generic instant t is
  s   s0   s0
1 t, t0 =  0  + 0.40k2 t, t0 + k3 k4 t, t0 (3.44)
E t0 E (28) E (28)

The rst term is the instantaneous elastic deformation of concrete, which is fully recovered when
the compressive stress is relieved. The second term is the deferred elastic deformation, which is also
reversible and may reach 40% of the instantaneous elastic deformation at 28-day curing based on
a coefcient 0.75 , k2(t,t0 ) 1 that depends on the load duration and can be assumed equal to
one for periods longer than 3 months. The third term is the irreversible deformation due to creep.
The creep strain is higher the younger the concrete is when loaded. It depends on the hygrometric
conditions, on the thickness of the element, and on the quality and compactness of the concrete, as
creep is caused by sliding of grains due to the viscous behaviour of the cement paste (Collepardi,
1980). The coefcients k3 and k4(t,t0 ) are dened by the applicable design standards.

Most bridge design standards lead to expressions for Equation 3.44 in the form 1(t, t0 ) = s0 /
E(t, t0 ), where E(t, t0 ) is a conventional long-term elastic modulus of concrete, and the analysis
of the total deformation becomes an elastic calculation with a conventional modulus.

3.5.3 Strand relaxation


The time-dependent behaviour of prestressed concrete is complicated by viscous phenomena in
prestressing steel known as steel relaxation. Relaxation occurs in a tensioned tendon of constant
length and causes loss of prestress.

As steel relaxation develops over a short period of time, in the past tendon re-tensioning a few days
after the initial stressing has often been adopted. Re-tensioning is expensive, and impossible when the
tendons are spliced with couplers or anchored at the match-cast surface of construction joints. With
the commercial availability of low-relaxation strand, tendon grouting after the initial tensioning has
become standard practice.

The evolution of steel relaxation can be expressed as


Dsp (t)
log = k5 + k6 log(t) (3.45)
sp,0
where Dsp(t) is the loss of prestress at time t, sp,0 is the initial tensile stress in the strand, and the
coefcients k5 and k6 depend on the properties of steel and are dened experimentally. In reality,
the length of the tendons is not constant over time because of the time-dependent shortening of
concrete, and the loss of prestress due to steel relaxation is less than the theoretical value. Some
design standards specify reduction coefcients to account for this effect.

190

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


Prestressed concrete bridges

3.5.4 Effects on bridge launching


In a continuous beam supported on sliding bearings, creep and shrinkage of concrete shorten the
beam. Neglecting the friction resistance of the bearings and the loss of prestress, shortening does
not modify the stress distribution during construction and in service.

The launch tendons extend for two or three deck segments to save on labour and anchorages, and
launch prestressing is therefore applied gradually to the deck. Integrative post-tensioning is applied
on launch completion, and therefore after a long curing time. The time-dependent shortening of the
deck is smaller than in bridges prestressed after a short curing time, and the time-dependent losses of
prestress are smaller as well.

In long launched bridges, the front segments of the deck reach their nal position many months
after casting, and the segments cast after them are progressively younger. Shrinkage, creep and
strand relaxation vary along the deck in relation to the age of the segments. The length of the
segments is progressively increased during deck construction to align the pier diaphragms with the
supports when the deck is placed on the permanent bearings on launch completion.

At the end of launch, creep of concrete does not alter the stress distribution signicantly. The
effects of creep can be assimilated as a reduction in the elastic modulus and, if the reduction is
almost uniform along the deck, the stresses are unaffected. The irreversible creep deections
developed by the deck during launch modify the theoretical distribution of the support reactions,
which is re-established by weighing and correcting the support reactions during deck positioning
on the permanent bearings to avoid long-term stress redistribution.

The situation is different during launching, because the support congurations of the deck
change continuously. Casting a new segment takes 12 weeks, and during that period the deck
develops irreversible creep deections related to that support conguration. New creep deec-
tions develop during the subsequent launch stops, the cylindrical deck geometry distorts, the
prole of the continuous beam becomes non-launchable, and displacing a non-launchable
prole on aligned launch bearings causes hyperstatic stresses similar to the effects of bearing
misalignment.

The effects of thermal gradients, geometry irregularities and misaligned launch bearings are too
irregular to cause signicant time-dependent effects. Axial launch prestressing causes negligible
bending, and the creep deections developed during the launch stoppages are therefore related
to self-weight only. When the launch step is as long as the span (one-span segments or shorter
long-line segments launched as a whole) and the spans have constant length, during construc-
tion of the new span the deck is subject to the same self-weight bending as in the previous
launch stop. Although shifted one span forward, the cross-sections are subject to the same
exural rotations and, as time goes on, the creep of the concrete makes a part of these rotations
irreversible.

Let us consider a bridge with regular span distribution, a launch nose with Ln /L = 0.6 and qn /q =
0.1, and full-span launch steps that end with a front PC cantilever as long as one-third of the
span. During each launch stoppage, negative bending in the front support section of the deck is
MB = 0.094qL2. If the end span of the bridge is 80% of the typical span, the deck section on the
front support during the launch stoppages will be subject to a moment MB = +0.047qL2 in the
nal position at the end of launching. Construction by incremental launching may take several

191

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


Bridge Launching

months, and a substantial portion of the exural rotation related to a large bending differential,
DM = 0.141qL2, becomes irreversible. The deck elevation at the end abutment and the rst pier
will not be correct at the end of launch, and the support reactions will have to be corrected during
insertion of the permanent bearings (AFGC, 1999).

The support reactions are typically corrected, which avoid bending redistribution. However, even
if the reactions were not corrected, the effects of bending redistribution would not be critical,
because launch prestressing is designed for controlled crack width during launch in the presence
of thermal gradients, geometry imperfections, misaligned launch bearings, midspan loads of pier
diaphragms, and time-dependent deck deections and ensures wide margins from the SLS tensile
stress limit in the nal sequence of lightened and stiffened sections on launch completion.
Additional edge compression results from the application of integrative post-tensioning.

The creep deections of the deck increase with the duration of the launch stoppages, which
depends on the duration of the casting cycle and on the number of casting cycles. When the bridge
is long and the bending redistribution is signicant during launch, the casting cell may be moved
closer to the abutment to avoid long front cantilevers during the launch stoppages, and the launch
nose may also be lengthened for earlier landing.

Creep deections in the front span progressively rotate the nosedeck joint and misalign the
launch nose. The launch bearings may be placed onto jacks to obtain the theoretical support
reaction on the launch nose at deection recovery and then during the course of launch. The
launch nose may be narrower than the deck soft in order to use hydraulic bearings only for the
nose. In this case the pier-cap gear includes two hydraulic bearings for the nose and two standard
launch bearings for the PC deck. The hydraulic bearings are moved to the next pier after arrival of
the PC deck.

In most launched bridges, the launch steps are half the span length, self-weight bending superim-
poses every two launch steps, and during casting of the segments most of the deck is subject to
exural rotations opposite in sign to those of the previous launch stop. The irreversible creep
rotations tend to cancel each other out, their envelope becomes irregular, and the cumulative effect
reduces signicantly. The effects of creep are also less signicant when the span distribution is
irregular.

As launch goes on, these geometry defects disturb the launch stresses. Disturbance is negligible in
the shortest bridges as their construction duration is too short for signicant time-dependent
effects to develop. In longer bridges, the creep deections are calculated under the assumption
of ageing linear viscoelastic material, by assuming that any stress increment generates an irrevers-
ible deformation that does not depend on the stress increments applied previously.

The analysis is repeated for each launch stop. The irreversible creep deections developed during
the stop are added to those developed during the preceding stops to obtain the new deformed
prole of the deck. The movement of the deformed prole on aligned launch bearings causes
hyperstatic launch stresses. The analysis is performed by modelling the deck as undeformed and
by applying to the support points of the continuous beam a vertical displacement corresponding to
the irreversible deck deection at those points. On launch completion, the launch nose is removed
from the model, and superimposed dead loads and polygonal integrative tendons are applied to
the model for long-term analysis of the time-dependent effects.

192

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


Prestressed concrete bridges

Figure 3.38 Upper edge axial stress in the Amiens Viaduct. (Reproduced with permission from ASCE)

20

Upper edge stresses: MPa

10

0
Front spans

The analysis of the hyperstatic effects of the irreversible creep deections requires laborious
calculations. Many structural analysis programs are able to analyse the time-dependent deections
of staged construction, but they are not written to apply cumulative deections to the support
points of the continuous beam as load conditions for the next steps in the analysis. During the
preliminary design, it is often sufcient to keep a margin of 0.40.6 MPa from the SLS tensile
stress limit to account for the effects of creep. Time-dependent launch analysis may also be
avoided by means of additional geometry imperfections applied to the model as instantaneous
load cases.

Figures 3.38 and 3.39 show the edge axial stresses in the three front spans of the Amiens Viaduct in
France. The stresses were calculated by taking creep of concrete and instantaneous prestress into
account (thicker line), and by disregarding creep and using the launch prestress at the end of

Figure 3.39 Lower edge axial stress in the Amiens Viaduct. (Reproduced with permission from ASCE)

20
Upper edge stresses: MPa

10

0
Front spans

193

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


Bridge Launching

Figure 3.40 Shimming of nose bottom flanges

launch (thinner line). The difference in the stress contours conrms the importance of the phenom-
enon. Although creep of concrete reaches 8590% of its nal value in 10 years, 30% is developed
in the rst month, 50% in 3 months and 7580% in 1 year. Therefore, because of the duration of
construction of a medium-length bridge most of the creep of concrete is mobilised.

Large creep deections may develop in the case of stoppage of work (Rosignoli, 1998a). The
deections developed during a 21-month stop in the launch of the Serio River Bridge in Italy
required corrective devices such as shimming of the nose bottom anges and integrative local
prestressing in order to complete the launch within the SLS stress limits specied by the design
standard (Figure 3.40).

3.6. Prestressing
In incrementally launched bridges, prestressing performs two primary functions. As with any
other bridge, prestressing resists the effects of self-weight, live load and other service loads in the
completed structure. In addition, it resists self-weight bending during launching. The bending
moments corresponding to these two functions have different distributions, and it is often not
practical to provide a single arrangement of tendons to perform both functions. Instead, two
separate post-tensioning concepts are generally provided, each adapted to the needs of one of the
two primary functions (Ontario, 2006).

The edge tensile stresses associated with incremental launching on short spans could often be
resisted by mild reinforcement. In most cases, launch prestressing is used to reduce the edge tensile
stress. Launch prestressing makes the cross-section lighter and more ductile and exible, controls
edge cracking, enhances the shear capacity of the webs, and reduces rebar congestion.

Launch prestressing is typically designed to avoid edge decompression in semi-permanent con-


ditions (self-weight bending and hyperstatic effects of creep deections in the deck). The effects
of thermal gradients, misaligned launch bearings and geometry imperfections are resisted with

194

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


Prestressed concrete bridges

reinforcement designed for control of crack width. Wind is rarely considered for the SLS launch
checks of a PC box girder. The ULS checks typically include factored prestressing, self-weight and
lateral wind.

Cyclic sign reversal of bending and shear during launch does not allow the use of undulated nal
tendons for launch prestressing, as their action would be inappropriate in the transient support
congurations that differ from the nal one. The resultant force of launch prestressing is kept cen-
troidal, to uniformly compress the deck cross-sections and resist tensile stresses indifferently at
both edges. The axial prestressing force is, therefore, higher the lower is the governing section
modulus.

The obvious solution to obtaining uniform centroidal prestressing would be pretensioning, which
is simple to attain using rectilinear strands. Pretensioning would avoid the cost of anchor beams,
tendon anchorages, ducts, grouting and skilled labour, and would also ensure better compaction
of the concrete and better strand corrosion protection. However, the average compressive stress
that would need to be applied to the cross-section is rather high, 57 MPa for full-span launching
and 23 MPa for launching on temporary piers, and would result in prestressing forces of tens
of mega-newtons. Applying such forces would require expensive anchor blocks, reaction beams
and lock systems for the deck, which is supported on low-friction bearings and can be pulled
backwards with much smaller forces. Therefore, launch prestressing is always obtained using
post-tensioned tendons, arranged either internally or externally to the cross-section.

Launch prestressing could be designed by enveloping launch stresses and nal service stresses so as
to apply the entire deck prestressing during segment construction. Uniform compression is not
very effective on launch completion, as it loads deck edges already compressed by permanent
bending and does not control the shear force at the supports. Launch prestressing is therefore
designed for the launch stresses, and more specialised prestressing systems are applied on launch
completion to prepare the deck for service conditions.

The cost of launch prestressing increases with the span length, and several solutions have been
attempted to reduce its cost on long spans. As a substantial amount of prestressing is available
from the beginning of launch, the rst idea was to shift the tendons on launch completion from
the centroidal position used during launch to an eccentric alignment, so as to use the same strands
for both the launch prestressing and the nal prestressing. The bridge launched over the Rio
Caroni in Venezuela (Baur et al., 1966) was actually prestressed in this way. Savings in material
were offset by labour costs and the duration and complexity of shifting the loaded tendons, and
this prestressing scheme was abandoned.

More specialised prestressing schemes were studied (Rosignoli, 1999a), consisting of permanent
launch tendons integrated on launch completion with eccentric cables. Thus launch prestressing
became a permanent component of deck prestressing, and not only because of the cost of its
removal. Axial prestressing increases the margins from edge decompression in the case of unfore-
seen events, such as support settlement or seismic response. It also improves the deck behaviour at
the counterexure points of permanent bending, covering the wide movements of the null points
produced by the presence or absence of live loads and thermal gradients.

The entity of the integrative eccentric post-tensioning depends on the loads applied to the deck on
launch completion, on their distribution between permanent and live loads, and on the moment of

195

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


Bridge Launching

inertia of the cross-section. Heavy railway bridges typically require substantial prestress integration
on launch completion because of the large service loads and the low span-to-depth ratio. Substantial
prestress integration is also required when the deck is launched on temporary piers that halve the
launch span.

As launch prestressing tends to increase the total cost of prestressing (Rosignoli, 1999b), the pre-
stressing schemes for launched bridges have been constantly improved to make the centroidal
component of nal prestressing more effective and less expensive. The optimum layout of the
launch tendons depends on the span length, the deck segmentation, the location of the casting cell
in relation to the launch abutment, the casting cycle, the launch staging in relation to casting of
segments, the length of the launch operations (half span versus full span) and the type of devices
used to launch the deck.

3.6.1 Launch prestressing


Longitudinal bending is negligible in the new deck segment supported on the extraction rails of the
casting cell. In most cases, prestressing is applied before launch to avoid shrinkage cracks and net
tensile stress during segment extraction from the casting cell, and to prepare the segment for the
exural stresses on the curing spans between casting cell and launch abutment, and ultimately on
the end span of the bridge.

The evolution of the incremental launching construction method has seen several schemes used for
launch prestressing (ASBI, 2008).

g permanent internal rectilinear tendons spliced by couplers at the construction joints


between subsequent deck segments
g temporary or permanent internal rectilinear tendons anchored in accessible blisters and
spliced by overlapping (the temporary tendons are typically in the top slab in the midspan
and in the bottom slab in the support regions)
g temporary or permanent external rectilinear tendons, either together with one of the
previous types or alone
g permanent draped tendons, either internal (parabolic) or external (polygonal), along with
temporary antagonist tendons that make the resultant prestressing force rectilinear in
most cases the resultant force is at the middle of the depth of the cross-section, and
additional temporary external rectilinear tendons are used to lift the resultant force to the
deck gravity axis.

When the integrative parabolic service tendons are internal they are located within the webs, and
the internal launch tendons are distributed across the two slabs in order to avoid interference.
Modern launched bridges use external tendons for integrative end-of-launch prestressing, and the
webslab nodes are available for powerful launch tendons that reduce the cost of launch pre-
stressing and can be spliced using couplers at the construction joints to avoid the cost and complexity
of anchor beams. In most cases, however, the launch tendons are located in the slabs.

When the curing area between the casting cell and the launch abutment is sufciently long, the
internal launch tendons cross two or three deck segments, and one-half or one-third of the tendons
is anchored at each construction joint. Longer launch tendons are less powerful because of their
number and because of symmetry considerations, and smaller tendons minimise the thickness of
the anchor beams, and simplify reinforcement and stress dispersal within the cross-section.

196

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


Prestressed concrete bridges

Figure 3.41 Precast anchor blocks in a prefabricated cage. (Reproduced with permission from ASCE)

The launch tendons are tensioned to full capacity and grouted, which requires adequate concrete
strength at the rear anchorages. Precast anchor blocks may be used in combination with cage
fabrication within the casting cell or with full cage prefabrication (Figure 3.41). Concrete defects
or honeycombs behind the anchor blocks can break the blocks during tensioning of the tendons
(Bernard-Gely and Calgaro, 1994), and the geometry of the joint surfaces and the rebar protrud-
ing from the joints are designed to facilitate concrete vibration. As an alternative, concrete around
the anchorages can be locally heated by means of thermal insulation, warm air or immersed elec-
trical resistance. Local heating permits a better continuity of reinforcement than does the use of
precast anchor blocks. Higher-strength concrete can also be used in the zones around the
anchorages.

3.6.1.1 Internal tendons spliced by couplers


After the initial experience gained in the shifting of tendons, launch prestressing was achieved
using short rectilinear tendons that were as long as the deck segments, and spliced with couplers
at each construction joint. The use of couplers offers the advantages of low permanent anchor
stresses, low friction losses of prestress in the rectilinear tendons, and simple forming and
reinforcement because there is no need for anchor beams or double blisters. The couplers,
however, also have several limitations:

g a high cost of prestressing due to the short tendon length, the cost of the couplers, and the
high demand for skilled labour; uncoilers and winches are also necessary to install
prefabricated tendons
g the impossibility of removing a portion of prestressing on launch completion

197

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


Bridge Launching

g the need to increase the thickness of the top and bottom slabs at the construction joints to
contain the couplers
g the reduced exural capacity of the joint region due to the area of the couplers
g the sharp application of full launch prestressing to fresh concrete.

The application of full launch prestressing may cause joint cracking because of elastic strains in the
concrete behind the couplers. The force applied by each anchorage deforms the concrete locally,
and with it the construction joint against which the next segment is match-cast. Tensioning the
newly added tendons relieves the anchor stresses at the couplers in the preceding segment, and
recovery of concrete strains compresses the joint locally and decompresses the zones further from
the anchorages, resulting in tensile cracks.

Unless launch prestressing is higher than 3.54 MPa, the zones of the joint section far from the
couplers should have a specic reinforcement to control cracking (Leonhardt, 1979; Rosignoli,
1999a). It is also expedient to use many small tendons to ensure a good distribution of the pre-
stressing force in the cross-section, which further complicates the joint geometry. For these
reasons, the evolution of this scheme of launch prestressing was rapid, and followed two different
directions.

g In a rst scheme, the launch tendons are coupled at each joint but tensioned alternately
every other joint, resulting in a more progressive application of prestress (Figure 3.42).
Every two-segment tendon has a coupler anchorage at the ends and individual strand
couplers in the middle. This prestressing scheme is expensive but complies with those design
standards that require that not more than 50% of the longitudinal tendons be coupled at
one section to control the axial stress anomalies at the joint. In the past, this scheme was
also applied by means of prestressing bars.
g In a second scheme, the couplers are staggered so as to prestress two deck segments with each
tendon and to halve the number of couplers (Figure 3.43). When the curing area is of
sufcient length, the launch tendons can be coupled every third joint to save on couplers and
to make application of launch prestress more progressive. If the strand is inserted into empty

Figure 3.42 Tendons coupled at each joint and stressed alternately at every other joint. (Reproduced
with permission from ASCE)

Launch direction

Segments I+3 I+2 I+1 I

Coupler anchorage on a tensioned tendon

Strand couplers

198

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


Prestressed concrete bridges

Figure 3.43 Staggered couplers. (Reproduced with permission from ASCE)

Launch direction

Segments I+3 I+2 I+1 I

Coupler anchorage

ducts after casting the segments, recesses are necessary at the couplers to anchor the strands.
Prefabricated tendons emerging from the rear joint simplify coupling but complicate cage
assembly in the casting cell, and make full cage prefabrication impossible (Figure 3.44).

The use of couplers exposes the process to the risk of detachment of the anchor plate from the
support plate when the new tendon is stressed. Short rectilinear tendons are subject to minimal
friction losses, anchor set and the rst time-dependent effects reduce the pull in the coupled

Figure 3.44 Two-segment prefabricated tendons emerging from the joint. (Reproduced with
permission from ASCE)

199

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


Bridge Launching

tendon, the new tendon is overstressed to compensate for anchor set, and the force transferred by
the new tendon may therefore be higher than the anchor load of the coupled tendon. The stress in
the strands is uneven and, if the anchor plate detaches from the support plate, plate rotation may
misalign and break both cables.

In spite of these limits, coupling the launch tendons may be the most effective solution for slender
box girders where access to the box cell is difcult, and for solid and voided slabs. Recommended
precautions for the use of couplers are to limit the number of tendons coupled in a section to not
more than one-half of the total number of tendons, to check the joint section for an ULS where the
force transferred by the couplers is reduced by one-third, and to avoid lap splices in the longitudi-
nal mild reinforcement for a certain distance on either side from the joint (AFGC, 1999).

3.6.1.2 Internal tendons spliced by overlapping


Pushing strands into empty ducts on the job site has become an attractive solution. The strand can
be inserted before, during or after casting the segment. The hydraulic pushing motor is installed in
a xed location and connected to the strand insertion point with a exible pipe. The strand pushers
provide adjustable speeds up to 8 m/s, and require minimal operating personnel and maintenance.
The strand coils are stored in a sheltered location and handled by the tower crane. These advan-
tages have made pushing the preferred method for strand installation.

Pulling an entire bundle of strands throughout the duct with a winch and a drive cable is mostly
used for U-tendons with tight loop anchorages. The strands are bundled close to the installation
point and protected for reasons of cleanliness and to avoid dragging extraneous materials into the
duct. The bundle of strands is inserted into the duct after casting of the concrete.

Inserting strands into empty ducts has made splicing of the launch tendons by overlapping a very
cost-effective solution. It has several advantages:

g Handling empty ducts instead of heavy prefabricated tendons reduces the risks to workers,
saves labour, speeds up cage assembly and ensures a better nal result (Rosignoli, 1998c).
g The launch tendons typically cross two or three deck segments. When the bridge has short
spans and adequate space is available behind the abutment, the launch tendons can be as
long as two spans. When the deck of a highway overpass is cast full-length behind the
abutment for launching in one operation, full-length launch tendons are anchored at the
ends of the deck. Long launch tendons reduce the labour demand, the cost of anchorages
and the prestress losses due to anchor set. The friction losses are relatively low in rectilinear
tendons and may be reduced further by stressing from both ends. Friction tests are easy to
perform using two stressing jacks and two load cells on the tendon.
g The absence of launch tendons emerging from the construction joint permits integral cage
prefabrication.
g The concrete curing time can be used to fabricate the tendons. This reduces the criticality of
steel cage assembly within the casting cell and improves labour rotation.
g The cost and complication of couplers are avoided.

In many PC launched bridges, the launch tendons have been distributed within the slabs to leave
the webs available for the parabolic tendons of integrative service prestressing. Several recent
launched bridges have used external polygonal tendons for integrative prestressing, and a few
powerful launch tendons have been placed at the webslab nodes and integrated with launch

200

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


Prestressed concrete bridges

tendons in the slabs. In prestressed composite box girders with steel corrugated-plate webs, launch
tendons in the slabs are integrated with external rectilinear launch tendons.

The launch tendons are anchored at the construction joints of the deck, which are at the span
quarters for half-span segments, and at the rear face of the pier diaphragm for full-span segments
and short segments cast using long-line techniques. In a box girder, the distance between the cross-
section centroid and the extrados is 3040% of the deck depth, and the number of launch tendons
in the top slab is, therefore, twice the number of tendons in the bottom slab to achieve a centroidal
prestressing force. When the launch tendons cross two or three segments, the number of tendons
typically increases. Long tendons are therefore relatively small, often 9T15, which facilitates the
distribution of their anchorages within thin concrete slabs. Smaller tendons are used when the
deck is launched over temporary piers. The slabs are thin, the ducts are often at the minimal dis-
tance from the surface allowed by the design standards, and the tendon design layout must there-
fore be respected carefully. The ducts are supported closely, and concrete segregation and
honeycombing must be avoided.

The front anchorages emerge within the box cell from anchor beams connecting the webs of the box
girder at the rear end of each segment. The anchor beams are as thin as possible, compatibly with
the dimensional and cover requirements of the anchorages, and are 24 m long to ensure sufcient
radius of curvature of the ducts and adequate tendon overlapping. When only a couple of launch
tendons are spliced, full-width anchor beams are replaced with double blisters at the webslab
nodes. The use of anchor beams at the centre of the slabs is not recommended (AFGC, 1999).

The rear anchorages may also emerge from double anchor beams to balance the load on the
anchor beams after tensioning of the rear set of tendons. The double anchor beams are designed
for the load imbalance prior to tensioning of the rear set of tendons. Rear anchorages in double
anchor beams are always accessible and permit relief of some of the tendons on launch completion
(Figure 3.41). This operation is labour intensive and is adopted only to correct the undesired
effects of launch prestressing. Recovered strand is hard to reuse, anchorages and ducts are lost,
the ducts must be grouted to avoid empty holes in the deck, and non-grouted launch tendons are
exposed to the risk of stress corrosion during launch.

When the rear anchorages are located within the construction joint, match-casting of the new
segment covers the anchorages, the launch tendons cannot be removed, and special reinforcement
is necessary because of the creep of the concrete. Unlike coupled tendons, the anchor forces do
not reduce over time. Creep deformations in the previous segment decompress the joint regions
in front of the anchorages, and local reinforcement is necessary to control joint cracking
(Leonhardt, 1979).

3.6.1.3 External rectilinear tendons


The bridge launched over the Rio Caroni in Venezuela did not use permanent launch prestressing.
Service prestress was obtained by varying the position of the launch tendons, which for this
purpose had to be external. Later, the use of permanent internal launch tendons integrated with
draped service tendons on launch completion became the standard approach. In some bridges,
however, intermediate solutions based on three families of tendons have been tested (Figure 3.45):

g internal rectilinear permanent tendons distributed in the two slabs so as to produce a


centroidal prestress force

201

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


Bridge Launching

Figure 3.45 Internal final launch tendons combined with external temporary tendons. (Reproduced
with permission from ASCE)

Temporary tendons Final tendons

g external rectilinear temporary tendons, also centroidal, to be relieved on launch completion


to reduce axial compression without hyperstatic effects
g internal integrative tendons fabricated and tensioned on launch completion (Figure 3.46).

The use of some external rectilinear temporary tendons for launch prestressing complicates the
inner forms and steel cage but offers many advantages, such as fewer embedded ducts and

Figure 3.46 Internal integrative tendons. (Reproduced with permission from ASCE)

Span tendons Cap tendons Parabolic tendons

202

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


Prestressed concrete bridges

grouting operations, simple tendon removal, and higher exural capacity of the cross-section due
to the absence of non-grouted ducts during launching (Breen and Naaman, 1990). Collars are used
to anchor the tendons to the inner surface of the box girder, and the bundles of strands are inserted
into sheet metal ducts or high-density polyethylene (HDPE) pipes for the protection of personnel
in case of strand breakage. Reusing the strand of the temporary tendons is rarely cost-effective
because of the numerous precautions that need to be taken.

The external launch tendons may be anchored in special double blisters or to the pier diaphragms
in order to reuse their anchorages for integrative external prestressing in the future. In this case, the
pier diaphragms must be cast before launching, and their weight increases the demand for launch
prestressing. The need for curing of the pier diaphragms may also increase the length of the casting
yard. When the deck is short and launched in one operation (single spans or short highway over-
passes), full-length external launch tendons may be anchored to the end diaphragms. Temporary
launch tendons are rarely used in railway bridges because of the powerful integrative prestressing
applied at the end of launching.

Using a large number of temporary launch tendons is rarely cost-effective, even when the bridge
owner allows reuse of the strand recovered from the external launch tendons for the integrative
end-of-launch prestressing. External launch tendons can be very useful in the front deck region,
as they permit design of the internal launch tendons for the launch stresses of the rear deck region.
Eccentric external launch tendons are also used to align the resultant prestressing force of
antagonist tendons with the deck gravity axis. External tendons rarely exceed 25% of the total
launch prestressing.

3.6.1.4 Antagonist tendons


Antagonist prestressing has been used on several bridges in recent years. The principle is to separate
the nal prestressing from the temporary one. The nal prestressing is the typical one for bridges cast
in-place or erected span-by-span using a self-launching gantry, with internal parabolic tendons or
external polygonal ones, sometimes complemented by internal cap and span tendons (Rosignoli,
2013). The temporary prestressing consists of tendons that compensate for the eccentricity of the nal
tendons during launch, until a centroidal prestress force is obtained. Temporary tendons can be:

g Short and rectilinear, placed inside or close to the bottom slab in the support regions and
to the top slab in the midspan regions. These tendons are spliced by overlapping with the
permanent tendons in anchor beams that, in the case of external integrative prestressing, are
reused as deviation diaphragms on launch completion. With this prestressing scheme, the
optimum length of the deck segments is one-third of the span for the anchor beams to be
located at the construction joints. This type of antagonist prestressing is rarely used
nowadays.
g Long and with a polygonal layout symmetrical with the nal tendon layout. Deviation
saddles are necessary also at the top slab, which complicates the use of a sliding form table
for two-phase casting.

In a conventional launched bridge with internal rectilinear launch tendons in the slabs, the launch
tendons represent about two-thirds of the nal prestressing, and the draped integrative end-of-
launch tendons are one-third of the total. When antagonist prestressing is used, launch prestressing
includes one-third of rectilinear tendons, one-third of permanent draped tendons, and one-third of
temporary tendons with antagonist layout. On launch completion, the antagonist tendons are

203

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


Bridge Launching

repositioned with the nal layout, so that two-thirds of the nal prestressing participates in the shear
reduction, compared with one-third of the rst case.

Axial prestressing during launch is 50% higher. High axial compression reduces the principal
tensile stress in the webs and leads to thin webs and efcient cross-sections. The need not to com-
promise this result with anomalies in the launch shear at the deviation points (especially when
internal parabolic tendons are compensated for by external polygonal tendons) complicates the
positioning of the deviation points and the cost of inner forming. Launch shear anomalies are
minimised when the draped tendons are all external on launch completion.

The use of antagonist tendons leads to optimised prestressing schemes during construction and in
service. Antagonist prestressing is an interesting solution also for the launch of solid or voided
slabs, where the standard prestressing scheme (two-thirds of launch tendons and one-third of
integrative end-of-launch tendons) is complicated by poor access to the anchorages and excessive
longitudinal compressive stress at the bottom edge of the support regions (AFGC, 1999).

The tendons are generally chosen for strength in order to reduce their number and avoid excessive
congestion within the box cell. At the end of launch, the antagonist tendons are released, and the
strand may be reused for the permanent tendons. For safe reuse of the strand, some precautions
are necessary during its temporary use:

g Temporary launch prestress should be less than 8590% of the SLS tensile stress limit in
the strand, to reduce the risk of local strand plasticisation.
g Strand sections notched by the anchor wedges must be eliminated. This requirement is
often met by limiting the stress level during the temporary use of the strands, as the
additional elongation at nal tensioning moves the notched zones out of the anchorage. If
necessary, the permanent tendons may also be designed shorter than the antagonist ones.
g The strands should be placed in perforated HDPE pipes to ensure safety of workers and for
visual control of corrosion.
g Only one tendon is dismantled at a time. The bundle of strands is pulled into a temporary
HDPE pipe, the end of the pipe is lifted close the anchorage of a nal tendon of the same
span, and the same winch and drive rope are used to pull the bundle of strands into the
nal duct.

With the progressive increase of labour costs and the reduced tensile stress required for strand reuse,
the cost-effectiveness of reusing the antagonist tendons should be evaluated on a case-by-case basis.

The rst antagonist launch tendons were used in the 1980s. The voided slabs of the Sathorn Viaduct
in Thailand were launched using permanent internal parabolic tendons and temporary external
antagonist tendons deviated at multiple points (Figure 3.47) (Capitanio, 1985). The antagonist
tendons emerged above the top slab to increase midspan eccentricity, and were deviated using steel
saddles supported over the webs. In the support regions, the antagonist tendons were contained
within the webs (Figure 3.48). Permanent and temporary tendons were anchored in the support
regions for launch stages as long as the span.

Antagonist prestressing was also used for the Amiens Viaduct (Godardt et al., 1989). The launch
tendons were entirely external and located as follows: two with a nal trapezoidal layout deviated
at the span quarters, two with a nal triangular layout deviated at midspan, four in antagonist

204

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


Prestressed concrete bridges

Figure 3.47 Launch prestressing of the Sathorn Bridge. (Reproduced with permission from ASCE)

A B

A B

Temporary tendons
Steel
deviator

Temporary tendons

Steel deviators
Final tendons

AA BB

Final tendons Temporary tendons

position with a symmetrical layout, and six rectilinear and close to the top slab to lift the resultant
prestressing force to the deck gravity axis. On launch completion, span by span, the four antag-
onist tendons were relieved and repositioned parallel to the four permanent polygonal tendons
(Figure 3.49).

3.6.1.5 Prestressing of temporary joints


Most launched bridges keep the continuous beam scheme used during launch permanently. In
some cases, however, it is necessary to create intermediate expansion joints (Seifried and
Wittfoht, 1979). These temporary joints are designed to transfer shear and bending during launch
and to facilitate their separation when the deck has reached its nal position.

The presence of end pier diaphragms in both deck sections facilitates anchoring of external
tendons or bars for temporary joint prestressing. External prestressing reduces the congestion
of internal tendons in the end regions of the deck, but the joint tendons can also be contained
in the slabs (Figure 3.50).

205

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


Bridge Launching

Figure 3.48 Antagonist tendons emerging above the top slab. (Reproduced with permission from ASCE)

Figure 3.49 Prestressing operations on launch completion of the Amiens Viaduct. (Reproduced with
permission from ASCE)

206

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


Prestressed concrete bridges

Figure 3.50 Temporary joint prestressing with internal tendons. (Reproduced with permission from ASCE)

Section AA

Shear keys

Launch tendons Joint prestressing


A

A Integrative service tendons

The joint prestressing is designed under the same criteria used for the current cross-sections of
the deck, and distributed so as to produce a centroidal force. The joint compression ensures the
margins from edge decompression and joint slippage established by the design standards for dry
joints. If necessary, the joint can be equipped with shear keys, which are designed using conven-
tional criteria for match-cast precast segments.

3.6.2 Design of launch prestressing


In a PC deck with constant cross-section, composed of many equal spans, and equipped with a
launch nose of adequate length and stiffness, the envelope of launch bending is like that shown
in Figure 2.73. The deck can be divided into two zones: a rear region, characterised by a repetitive
course of launch bending and shear, and therefore by constant envelopes; and a front region,
characterised by peaks of bending and shear. Launch prestressing is typically designed for the rear
region and adapted to the higher demand of the front region.

Design standards and project-specic design criteria guide in the choice between three approaches
to the design of launch prestressing:

g No edge decompression for all SLS load combinations. This approach is specied only for
precast segmental decks with epoxy joints and for temporary prestressing of expansion
joints. Some design standards require a minimum compressive stress of dry joints.

207

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


Bridge Launching

g No edge decompression for semi-permanent SLS load combinations (self-weight bending


and hyperstatic effects of creep deck deections). Thermal gradients, bearing misalignment
and geometry imperfections can decompress the edges, and longitudinal reinforcement
within the segments and throughout the joints controls crack width. The crack width may
be calculated by enveloping the effects of full thermal gradients, full bearing misalignment,
full geometry imperfections, and a 50% combinations of these.
g Partial launch prestressing, allowing edge tensile stresses and reinforcement design for
controlled-width cracking independent of the SLS load combinations.

The most cost-effective approach for PC decks with through reinforcement in the joints consists of
allowing 1.52 MPa edge tensile stress during launch to be resisted by longitudinal reinforcement.
Reinforcement is designed for the total tensile force in the decompressed region of the cross-
section and a reduced tensile stress for control of crack width. It is important to remember that
launching is a transient construction phase, and partial prestressing during launch may be com-
bined with total prestressing in service if necessary.

In service conditions, self-weight, superimposed dead load and 1030% of the live load are
typically carried without edge decompression, and the remainder of the live load is carried with
decompressed edges. The edge tensile stress is not limited, and crack width is controlled by longi-
tudinal reinforcement according to the aggressiveness of the environment. Total prestressing may
also be specied for service conditions, which is achieved using more powerful end-of-launch
integrative tendons. The ULS checks are performed using the criteria specied in the design standards
for construction and service stages.

3.6.2.1 Rear deck region


In a cross-section subject to a centroidal prestress force F, the effects of a bending moment M may be
evaluated by introducing an eccentricity e = M/F to the application point of the force. In the rear
region of the deck, each cross-section is subject to bending moments that vary from Mmax to
Mmin . To avoid edge decompression, the axial prestressing must be such that the maximum upward
eccentricity eu = Mmax /F and the maximum downward eccentricity ed = Mmin /F maintain the
force inside the central core of the cross-section, the depth of which hc is dened by Equation 3.2:

eu + ed rfH (3.46)

This condition denes the minimum level of axial launch prestressing for the case of total pre-
stressing (no edge decompression for all SLS load combinations). Almost all launched bridges are
cast in-place, continuous reinforcement is present in the joints, and launch prestressing may
therefore be designed for no edge decompression under semi-permanent SLS load combinations
(self-weight and hyperstatic effects of creep deck deections). As no cross-section has a central
core perfectly dimensioned for Mmax and Mmin , decompression of one edge governs the design
of launch prestressing. Design standards specify SLS limits for the compressive stress sc and the
tensile stress st based on the age and strength of the concrete, and the edge stresses must fall
within those limits:

F M
st , + zu , s c (3.47)
A I

F M
st , zl , sc (3.48)
A I

208

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


Prestressed concrete bridges

The design governing condition is sought at both edges based on their distance zu (upper edge) and
zl (lower edge) from the centre of gravity of the cross-section:
Mmin F M
st zu , , sc max zu (3.49)
I A I
Mmax F M
st + zl , , sc + min zl (3.50)
I A I
These two equations can be rewritten to determine the minimum moment of inertia of the cross-
section:
H   
I max Mmin , Mmax (3.51)
sc + st

The centre of gravity of a box girder or a ribbed slab with a double-T section is typically well above
the middle of the depth of the cross-section. During launch, the peak tensile stress is often reached
at the lower edge of the midspan sections because of the small section modulus, even if the peak
positive bending Mmax is smaller, in absolute value, than the peak negative bending at the supports
Mmin . Therefore, axial prestressing is typically designed for the midspan tensile stress.

When the same cross-section is a support section, the upper edge may be far from decompression
because of the high section modulus, and the lower edge may be close to the SLS compression
stress limit, as axial launch prestressing adds to the high compressive stress generated by the high
Mmin and small section modulus. For the SLS tensile stress limit to be reached at the upper edge of
the support section and at the lower edge of the midspan section,
F Mmin F M
st = + zu = max zl (3.52)
A I A I
This equation can be rewritten to determine the optimal location of the centre of gravity of the
cross-section:
 
zl Mmin 
= (3.53)
zu Mmax
For self-weight bending, the optimal cross-section has the centroid at a distance from its upper
edge equal to one-third of its depth, Mmin being twice Mmax . Thermal gradients affect Mmax
more than Mmin , while misaligned launch bearings and the hyperstatic effects of creep deck
deections typically affect Mmin more than Mmax . The optimal location of the centroid is, there-
fore, 0.33 , zu /H , 0.42.

This condition is rarely satised because the wide top slab of a PC box girder draws the centroid
upward, and launch prestressing is therefore designed so as not to exceed the SLS tensile stress
limit at the governing edge. The number and power of the launch tendons are dened considering
tendon symmetry requirements, the number of deck segments crossed by the tendons, and the
need to use small anchorages to limit the thickness of the slabs. The launch tendons are particu-
larly small when temporary piers are used for launching. As a result of so many technological
requirements, the prestressing force is often higher than the minimum requirement.

Internal launch tendons are lodged within both slabs in a box girder, and in the top slab and at the
bottom of the webs in a ribbed slab. The same number of strands is used for the launch tendons in

209

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


Bridge Launching

order to use one stressing jack. The tendons are distributed such that the prestressing forces in the
top and bottom slabs are inversely proportional to their distance from the centre of gravity of
the cross-section. An even number of tendons is used in both slabs for reasons of symmetry.
When the top slab tendons are a multiple of three and the bottom slab tendons are not, the top
slab tendons may cross three deck segments, and the bottom slab tendons may cross two.
Because of the higher location of the centre of gravity, a ribbed slab with double-T section
typically requires more launch prestressing than does a box girder.

When external tendons are used for integrative service prestressing, the webslab nodes are
available for launch prestressing. Powerful launch tendons at the webslab nodes reduce the cost
of launch prestressing, and may be coupled at the construction joints to avoid the cost and com-
plexity of the anchor beams. Less powerful launch tendons in the slabs are used to lift the resultant
prestress force to the cross-sectional centroid level.

Launch prestressing is rarely centred perfectly throughout the length of the bridge because a thicker
bottom slab in the support regions lowers the centre of gravity of the cross-section. When prestress
eccentricity is constant, the hyperstatic effects nullify most of the isostatic effects of prestress eccen-
tricity in the rear deck region. The hyperstatic effects may be more pronounced in the front deck
region, and are sometimes considered in the design (Ramakrishna and Sankaralingam, 2000).

The possibility of applying additional prestressing during construction is rarely considered in a


launched bridge. Provisions for additional prestressing are frequently made in the balanced can-
tilever bridges because of the complex alignment of the cantilever tendons and the large number of
construction joints (Rosignoli, 2013). Launch prestressing is geometrically much simpler, and the
number of construction joints is much smaller in a launched bridge.

3.6.2.2 Front deck region


In the front region of the deck, Mmax and Mmin depend on the nosedeck interaction. The use of a
second-hand launch nose often involves some compromise, but even a custom-designed nose can-
not prevent negative bending and shear at the front support from being higher than at the rear
supports. The launch prestress used in the rear deck region must therefore be increased, and the
thickness of the bottom slab is often increased as well. When higher shear suggests the use of thicker
webs, vertical prestressing in the critical web regions is typically less expensive than a special set of
internal forms. The midspan moment may also require additional launch prestressing.

Internal local integrative tendons (in the top slab in the critical negative bending region, and in the
bottom slab in most of the front span) are used rarely, because of tendon congestion in the slabs.
The hyperstatic moment of local prestressing may also cause tensile stresses in the deck regions
behind the anchorages. The integrative tendons are mostly external and rectilinear. A few perma-
nent draped service tendons may be tensioned before launch in the two front spans in combination
with temporary antagonist tendons that compensate for their eccentricity (AFGC, 1999).

Additional temporary external rectilinear prestress Fa , with eccentricity ea positive if above the
deck gravity axis, is designed to meet the SLS edge stress limits in the design-governing launch
conditions (Rosignoli, 1998b). Neglecting the hyperstatic effects, at the front support it must be

F + Fa Fa ea Mmin
+ zu st (3.54)
A I

210

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


Prestressed concrete bridges

F + Fa Fa ea Mmin
zl sc (3.55)
A I
and for peak positive bending
F + Fa Fa ea + Mmax
+ zu s c (3.56)
A I
F + Fa Fa ea + Mmax
zl st (3.57)
A I
Additional launch prestressing may also be necessary at the rear end of the deck, especially when
Eberspacher launchers are used to move the deck. The front support of the curing area is far from
the launch abutment in order to increase the deck support reaction at the launcher, and this dis-
tance can become critical in the cantilever conguration at the end of launch.

Concrete is partially cured in the last segments of the deck. Instead of locally increasing the launch
prestress, it is preferable to introduce an auxiliary support between the front support of the curing
area and the launchers. This support is kept disengaged during most of launch by not inserting the
Neoon pads. The reduction in the support reaction at the launchers in the last launch cycles is
compensated for by a rear nose that anchors drawbars or strands pulled by the launchers
(Figure 2.36).

3.6.3 Service prestressing


On launch completion, the sequence of midspan and pier sections matches the demand of self-
weight bending and shear. Launch prestressing tends to be excessive for self-weight only, as a
better load distribution reduces bending and shear with respect to the launch envelopes, and
cross-sections designed for the nal load demand further reduce edge stresses and principal stresses
in the webs. Launch prestressing and web reinforcement are also designed for the effects of bearing
misalignment, geometry imperfections and creep deck deections, which are minimised during
weighting and correction of the support reactions for deck placement on the permanent bearings.
These margins, however, are typically insufcient to cover the service stresses, and launch pre-
stressing must be integrated to resist superimposed dead load and live loading.

The integrative tendons are draped so as to maximise the eccentricity of the resultant force. The
effectiveness of the correction depends on the quantity of axial prestressing in the deck, and ulti-
mately on the launch technique. When temporary piers are used to halve the launch spans, launch
prestressing is relatively low, the correction to apply at the end of launch is substantial (it includes
the additional self-weight bending from spans of double length) and the nal prestressing scheme
is more efcient. Final prestressing is also more efcient in railway bridges because of the high
design live load. Full-span launching of highway bridges requires many launch tendons, only a few
integrative tendons are necessary at the end of launch, and the efciency of nal prestress cannot
improve substantially, unless some launch tendons are relieved.

The efciency of parabolic prestressing in compensating permanent loads may be evaluated using the
equivalent vertical load ppar = Fpar /Rt , where Fpar is the parabolic prestressing force and Rt is the ver-
tical radius of the equivalent tendon (Cestelli-Guidi, 1987). The equivalent load balances a part of the
permanent loads (self-weight q and superimposed dead load psdl ), and the degree of compensation
ppar
b= (3.58)
q + psdl

211

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


Bridge Launching

is a good indicator of the exural stress distribution in the deck. Although it would be convenient
to reach b = 0.70.8 to limit the long-term midspan deections, b 0.6 is a reasonable target for
full-span launched bridges because of the substantial long-term compression provided by the
launch tendons in service.

The integrative end-of-launch tendons can be distributed over the cross-section in several ways,
according to the total force needed: with a parabolic layout within the webs, with an external
polygonal layout deviated by diaphragms or saddles, or with an external parabolic layout in con-
tact with the internal surface of the webs. Internal tendons are affected by high friction losses, are
not replaceable and increase the web thickness. Therefore, several recent launched bridges were
designed with external polygonal tendons deviated by diaphragms and designed for future
replacement, in spite of the smaller eccentricity achievable within the box cell. Powerful tendons
are used to reduce the cost of integrative prestressing and the tendon congestion within the box
cell. The tendons cross two or three spans and are tensioned from both ends. The pier diaphragms
have a long curing time on launch completion and can accommodate high anchor forces.

3.6.3.1 Internal integrative tendons


The internal integrative tendons include parabolic tendons within the webs, cap tendons in the top
slab in the support regions, and span tendons in the bottom slab at midspan.

The parabolic tendons rarely extend over several spans to limit the friction losses. In most cases the
tendons are limited to only one span and are anchored in single blisters at the counter-exure
points of the adjacent spans so that they overlap in the support regions with the parabolic tendons
of the adjacent span. This layout doubles the effect of the integrative tendons in the control of
shear and negative bending. If double stress correction is not necessary, only one double blister
is used per span (Figure 3.51).

Figure 3.51 Double blister for internal integrative tendons. (Reproduced with permission from ASCE)

Longitudinal section Section BB


B

A
A

Section AA

212

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


Prestressed concrete bridges

Splicing by overlapping requires tendon crossing and anchor beams or double blisters, which
increases the friction losses and the cost of parabolic prestressing, and complicates forming.
Some design standards also require 150 mm concrete cover to empty post-tensioning ducts at the
underside of the webs so as to not disturb the dispersal of the launch support reactions within the
webs, which reduces the eccentricity of the internal tendons. For all these reasons, the parabolic
tendons are often designed for control of shear, and additional prestressing demand is met using
local rectilinear tendons.

Cap tendons in the top slab are difcult to tension with overhead jacks but are rarely indispensa-
ble, as parabolic tendons combined with axial launch prestressing are often sufcient to control of
negative bending. Span tendons in the bottom slab are easier to install and structurally more
efcient because of the large eccentricity from the deck gravity axis. Span tendons are frequently
used in the end spans of the bridge, especially when they have the same length as the interior spans
and positive bending is therefore higher. The hyperstatic effects of the span tendons can help in
reducing the need for cap tendons.

The integrative tendons are fabricated throughout the length of the bridge during casting of the
last segments, and are tensioned from both anchorages on launch completion. Anchorages and
plastic ducts are embedded in the segments during construction, and are pressure tested for
watertightness prior to pouring of concrete. HDPE pipes or rubber mandrels as long as the
segment are inserted in the ducts prior to casting to minimise the risk of occlusion due to grout
spillage or squashing, although the pipes may be difcult to recover if they get stuck in
the duct. Torpedoes may be used to control duct squashing, and video-camera inspections
may be performed for additional risk mitigation. The ducts are washed prior to strand
insertion.

3.6.3.2 External integrative tendons


External integrative tendons are used more and more frequently because of the high quality and
replaceability of external tendons, the cost savings deriving from simpler web reinforcement, and
the high quality of monolithic webs. Without the clearance requirements of internal parabolic
tendons, the webs can be designed for control of cracking and the SLS principal compressive stress
limits specied in the design standards. Shear capacity is calculated for the full web thickness
because no internal ducts weaken the cross-section, and thinner webs reduce the weight of the deck
and increase the exural efciency of the cross-section.

In the absence of parabolic tendons within the webs, launch prestressing can be designed with a
few powerful tendons placed at the nodes of the cross-section. This reduces the thickness of both
slabs, and splicing the launch tendons with couplers avoids the need for anchor beams and double
blisters. Concrete in monolithic webs is stronger than concrete containing many ducts, which
disturb the stress distribution and complicate vibration. By eliminating internal tendons and their
supports, cage assembly is faster, and the density of reinforcement, the demand for skilled labour
and the effects of poor workmanship reduce.

Low friction losses lead to high tensile stress in the strand and high efciency of prestressing. Low
friction losses also allow the use of tendons as long as several spans, which reduces the number of
anchorages and the labour costs. In most launched bridges, integrative prestressing is introduced
on launch completion, and tendon length does not affect deck segmentation. The long curing time
of concrete also permits the use of powerful tendons.

213

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


Bridge Launching

External tendons are unaffected by concrete cracking and provide a more durable solution when
the bridge is designed for partial prestressing. Maintenance costs are also lower, as the tendons can
be inspected easily and repaired or replaced while keeping the bridge in service.

The anchor stresses of external tendons are higher than those of internal tendons due to the
absence of frictional load transfer and the use of powerful tendons to reduce tendon congestion
within the box cell. The tendons are typically anchored to the pier diaphragms and deviated using
saddles and diaphragms. The web thickness depends on the shear force in the support regions of
the deck, and therefore on the longitudinal distribution of tendon deviators (Figure 3.52).

Figure 3.52 Shear reduction in the deck. (Reproduced with permission from ASCE)

(a) Two diaphragm layout

(b) Fan-shaped layout

(c) Parabolic layout

(b) and (c)

(a)

Shear force reduction

214

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


Prestressed concrete bridges

Fan-shaped tendons from the pier diaphragms lead to a shear reduction close to the optimum
triangular one, and to a tendon layout close to the funicular diagram of the permanent loads.
Four or six deviation points in each span may lead to very thin webs and, as thin webs require
higher launch prestressing, the deviation diaphragms can also be used to deviate or to anchor
antagonist tendons. As an alternative to fan-shaped tendons, using the same polygonal layout for
all of the tendons reduces the angular deviation at the saddles closer to the pier, without modifying
the average angular deviation, and reduces interference and geometry errors. These deviators,
however, are typically heavier.

The weight of the deviators suggests reducing their number to two per span, and the tendon layout
becomes trapezoidal. When the deck segments are as long as the span or one-third of the span, the
deviation points are placed at the thirds of the span, which is the optimum location for bending.
The optimum layout for shear requires tendon deviation at the quarters of the span, and this is
the typical solution with half-span segments. Two deviators at the span quarters may also be
combined with a midspan deviator for better combined control of bending and shear.

3.6.3.3 Transverse prestressing


Transverse prestressing of the top slab is typically used in PC and composite decks wider than
15 m. It is denitely advantageous for decks wider than 1718 m, as transverse live load bending
increases with the square of the deck width. Prestress losses due to anchor set are less critical with
longer tendons, and the cost of anchorages is less prohibitive.

Above certain span lengths, the need for self-weight reduction makes transverse prestressing very
useful, as it results in a thinner deck slab. Slab deections, cracking and fatigue in reinforcement
are better controlled. For the same reasons, transverse prestressing is frequently used in the con-
crete slab of wide single-cell composite box girders based on a steel U-girder. Composite grillages
of steel I-girders rarely require deck prestressing because of the shorter distance between the main
girders and the frequent use of substringers as additional support lines for the concrete slab.

The level of prestressing can be chosen freely. Total prestressing is superuous and difcult to
achieve, as live load bending is much higher than self-weight bending. Rebar grids limit the vertical
eccentricity of the transverse tendons, especially in the slab regions distant from the webs. Partial
prestressing is based on closely spaced (0.51.0 m) light tendons (from 2T15 to 4T15) designed for
the transverse bending due to permanent loads and 3040% of live loads. Small anchorages do not
increase the slab thickness, and the stressing anchorages may be alternated with the dead ones
(Figure 3.53).

Figure 3.53 Alternated anchorage of transverse tendons. (Reproduced with permission from ASCE)

Stressing anchorage

Dead anchorage

215

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


Bridge Launching

Figure 3.54 Web prestressing with loop strand tendons. (Reproduced with permission from ASCE)

0.8 0.8

When the deck is designed for partial longitudinal prestressing, slab cracking in the support
regions of the continuous beam reduces the corrosion protection of the transverse tendons, and
the use of waterproof plastic ducts is therefore specied. With the advances in technology, plastic
ducts are more and more frequently used for the longitudinal tendons also.

3.6.3.4 Vertical prestressing


Axial launch prestressing cannot reduce the shear force in the webs, and each cross-section of the
deck resists full shear stresses during launch. The web thickness is designed to maintain the SLS
principal compressive stress within the limits specied by the design standards in relation to the
age and strength of concrete, and shear reinforcement is designed to provide additional shear
capacity and to control the width of the cracks caused by the principal tensile stress.

If the principal tensile stress exceeds the modulus of rupture of concrete only in localised regions of
the deck, the use of vertical prestressing may be a cost-effective solution (Belluzzi, 1988).
Prestressing bars or loop strand tendons anchored on the top slab are used to reduce web
reinforcement without modications of the cross-section geometry (Figure 3.54).

Vertical prestressing modies the principal stresses in the webs, reducing or eliminating the principal
tensile stress, and avoiding cracking. For vertical prestressing to be effective the webs must be deep,
which limits its use to long-span bridges. Some design standards specify that the tangential stress
(and consequently the principal tensile stress in the webs) must be calculated by neglecting the duct
diameter. In these cases, the use of vertical prestressing may be detrimental (Rosignoli, 1998e).

Vertical bars and loop tendons are short and expensive to install, and their effect is irregular because
of the anchor set. For these reasons, vertical web prestressing is used only to cover the peaks of the
principal tensile stress, while the remainder is resisted with conventional reinforcement. The use of
precast web panels vertically prestressed with adherent pre-tensioned strands could result in an
optimum state of stress in the webs, as well as labour savings and faster construction.

3.6.3.5 Recent trends


Although the use of external tendons in launched bridges is relatively recent, the particular
requirements of this construction method have led to several innovative solutions.

The rst applications developed the concept of mixed prestressing: rectilinear internal tendons in
the correct position for permanent bending, temporary rectilinear internal antagonist tendons

216

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


Prestressed concrete bridges

Figure 3.55 Final rectilinear internal tendons, temporary rectilinear internal antagonist tendons
(dotted), and polygonal external tendons. (Reproduced with permission from ASCE)

During construction

In service

removed on launch completion, and polygonal external tendons fabricated on launch completion
(Figure 3.55). Compared with the conventional tendon layout (launch prestressing in the slabs
and integrative parabolic tendons in the webs) this scheme results in thinner slabs and webs.
The disadvantages include the short length of the launch tendons, the need to grout the ducts
of the temporary tendons after their removal, and the risk of stress corrosion of the strand in
non-grouted ducts.

Other applications were based on permanent internal launch tendons integrated with rectilinear
external tendons during launch, and corrected on launch completion with rectilinear internal cap
tendons and span tendons, and with polygonal external tendons designed for shear correction
(Figure 3.56). In both cases, launch prestressing is obtained by superimposing several groups of
tendons according to deck segmentation. This makes it possible to relieve and remove some ten-
dons on launch completion. Long rectilinear external launch tendons are easy to inspect and to
relieve on launch completion, and no additional ducts are to be grouted in the slabs.

Other prestressing schemes derive from the concept of antagonist prestressing. In the basic scheme
(Figure 3.57), permanent parabolic internal tendons designed to balance most of the self-weight
shear are tensioned before launch along with cap and span tendons. The eccentricity of prestress
is compensated for by antagonist polygonal tendons and reduced by additional rectilinear
tendons, all external and temporary. On launch completion, the temporary launch tendons are
relieved, and prestressing is completed with polygonal tendons and a few span tendons, all external
and designed for the rest of self-weight bending and shear and the live loads. The parabolic
internal tendons diminish tendon congestion within the box cell. When internal parabolic tendons
are used for the integrative end-of-launch prestressing, the use of a few temporary antagonist
tendons in the rst two spans may solve the need for additional launch prestressing in the front
deck region.

217

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


Bridge Launching

Figure 3.56 Final launch tendons integrated with temporary launch tendons, and corrected with internal
tendons and polygonal external tendons on launch completion. (Reproduced with permission from ASCE)

Final internal tendons tensioned before launch

Temporary external tendons tensioned before launch

Final internal tendons tensioned on launch completion

Final external tendons tensioned on launch completion

A few launched bridges were designed with totally external prestressing (Combault et al., 1987).
Some polygonal tendons were tensioned in their nal position before launch, and they were com-
pensated by antagonist tendons and integrated by additional rectilinear tendons, all temporary
and external. The temporary tendons were relieved and repositioned on launch completion.

Typically, however, some rectilinear tendons are kept inside the concrete to reduce the congestion
of external tendons within the box cell (Figure 3.58), and to increase the ULS cross-sectional
capacity. In this case, prestressing may be designed as in Figure 3.59:

g powerful rectilinear internal launch tendons at the webslab nodes of the cross-section,
tensioned during segment construction and spliced by couplers or by overlapping in double
blisters
g one half of the polygonal external tendons tensioned during construction in the nal layout
g the other half used temporarily in an antagonist layout to be relieved on launch completion
and repositioned, span by span, in the nal layout

218

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


Prestressed concrete bridges

Figure 3.57 Final internal tendons compensated by antagonist tendons during launch and completed
on launch completion. (Reproduced with permission from ASCE)

Final tendons stressed before launch

Temporary tendons stressed before launch

Final tendons stressed at the end of launch

g temporary rectilinear external launch tendons that lift the resultant prestressing force to the
deck gravity axis
g a few internal span tendons in the bottom slab of the end spans of the bridge.
Removing several launch tendons reduces the residual launch prestressing, the polygonal tendons
are designed for most of the permanent loads, and this leads to very efcient permanent prestressing

219

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


Bridge Launching

Figure 3.58 External launch prestressing in the Charix Viaduct. (Reproduced with permission from
ASCE)

schemes. The high axial compression during launch leads to thin webs and high exural efciency of
the cross-section, and this further reduces the demand for launch and permanent prestressing.

3.7. Reinforcement
The reinforcement of a PC box girder built by incremental launching is designed using con-
ventional criteria. Two rebar grids are used in the top slab to resist longitudinal and transverse
bending due to permanent loads, live loads, thermal gradients and cross-sectional distortion.
Two rebar grids are also used in the bottom slab to resist transverse bending and to facilitate dis-
persal of the launch support reactions within the webs with local strut-and-tie mechanisms at the
bottom webslab nodes. Transverse reinforcement in the slabs also provides longitudinal shear
capacity at the webslab nodes for interface shear transfer of launch prestressing, as the launch
tendons are mostly located within the slabs. Two rebar grids in the webs provide additional shear
capacity and resist the SLS principal tensile stress generated by longitudinal shear, transverse
bending, thermal gradients and differential shrinkage, and control crack width in both directions.

Launch prestressing is typically designed for no edge decompression for semi-permanent SLS load
combinations and for controlled-width cracking for rare SLS load combinations. Longitudinal
reinforcement in the slabs is mostly designed for control of crack width. Several design standards
specify that the total tensile force in the decompressed region of the cross-section must be resisted
with reinforcement working at notional stress levels, which reduces the interest in the use of

220

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


Prestressed concrete bridges

Figure 3.59 Powerful rectilinear internal tendons at the webslab nodes reduce tendon congestion and
increase the ultimate flexural capacity. (Reproduced with permission from ASCE)

Final internal tendons stressed before launch

Temporary launch tendons

Final external tendons

Temporary antagonist external tendons

External tendons stressed after launch

high-grade steel. Typical reinforcement ratios for the tensile zone are 0.8% for box girders and
0.4% for ribbed slabs (AFGC, 1999).

Local reinforcement is often necessary to resist the thrust force. Eberspacher launchers apply loca-
lised longitudinal shear forces to the deck soft, which may locally exceed the tensile capacity of
concrete. When rear thrust beams or through pins are used for launching, it is necessary to resist
localised loads and to check the webslab nodes for interface shear transfer of the thrust force.

Transverse reinforcement includes straight and bent bars designed for the casting phases of the
cross-section. Bent bars are more expensive but are often necessary to stiffen the cage, as they act
as spacers of the bar layers.

221

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


Bridge Launching

Finally, it is a good practice to distribute wire mesh on the external surface of the deck to stiffen the
prefabricated cage during transfer and to control concrete cracking due to shrinkage and thermal gra-
dients. The area of this additional reinforcement may be about 0.1% of the deck cross-section, and
about 0.3% in the bottom slab, which is subject to the highest stresses during launch and in service.

Adequate stiffness of the steel cage is necessary for cage transfer from the rebar jig to the casting
cell and during casting and vibration, to prevent movements that could create cavities and honey-
combs around bars and prestressing ducts. Such cavities would permit corrosive agents to come
into contact rapidly with a large steel surface and accelerate depassivation and corrosion.

REFERENCES
AASHTO (American Association of State Highway and Transportation Ofcials) (2014) LRFD
Bridge Design Specications. AASHTO, Washington, DC, USA.
AFGC (Association Francaise de Genie Civil) (1999) Guide des Ponts Pousses. Presses de
lecole nationale des ponts et chaussees, Paris, France, p. 240.
ASBI (American Segmental Bridge Institute) (2008) Construction Practices Handbook for
Concrete Segmental and Cable-Supported Bridges, 2nd edn. ASBI, Buda, TX, USA.
Baur W, Leonhardt F and Trah W (1966) Brucke uber den Rio Caroni, Venezuela. Beton und
Stahlbetonbau 61(2): 2538.
Belluzzi O (1988) Scienza delle Costruzioni. Zanichelli Editore, Bologna, Italy.
Bennett MV and Taylor AJ (2002) Woronora River Bridge, Sydney. Structural Engineering
International 12(1): 2831.
Bernard-Gely A and Calgaro JA (1994) Conception des Ponts. Presses de lecole nationale des
ponts et chaussees, Paris, France.
Breen J and Naaman A (1990) External Prestressing in Bridges. Special Publication 120.
American Concrete Institute, Detroit, MI, USA.
Calgaro JA and Virlogeux M (1994) Projet et Construction des Ponts: Analyse Structural des
Tabliers des Ponts. Presses de lecole nationale de ponts et chaussees, Paris, France.
Capitanio S (1985) The King Taskin Bridge in Bangkok. LIndustria Italiana del Cemento,
September.
CEB/FIP (Comite Euro-International du Beton/Federation International de la Precontrainte)
(1993) CEB/FIP Model Code 1990. Thomas Telford, London, UK.
CEB/FIP (1994) Application of High Performance Concrete. Bulletin dInformation 222.
Federation Internationale du Beton, Lausanne, Switzerland.
Cestelli-Guidi C (1987) Cemento Armato Precompresso. Hoepli Editore, Milan, Italy.
Cezard C and Servant CI (1988) Charix Viaduct. Contribution of the French Group. AIPC
IABSE Congress, Helsinki. AFPC Publications, Zurich, Switzerland.
Collepardi M (1980) Scienza e Tecnologia del Calcestruzzo. Hoepli Editore, Milan, Italy.
Combault J, Leveille A, Neron P and Thibonnet JL (1987) Incrementally launched bridges with
total external prestressing. Contribution of the French Group. AIPCIABSE Congress, Paris,
France.
Crain RC (1997) Simulation using GPSS/H. Proceedings of the 1997 Winter Simulation
Conference, Atlanta, GA, USA, pp. 567573.
Cremer J, Counasse C and Delforno J (2003) The Sart Canal Bridge, Houdeng-Aimeries,
Belgium. IABSE Structural Engineering International (13)1: 1922.
Dezi L, Menditto G and Rosignoli M (1982) In tema di predimensionamento dei ponti a
cassone in calcestruzzo armato ed in calcestruzzo armato precompresso. LIndustria Italiana
del Cemento 6: 549570.

222

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


Prestressed concrete bridges

Frizzi D and Giovannini B (1977) Ponte sul Po a Casale Monferrato per lAutostrada dei
Trafori. LIndustria Italiana del Cemento, April.
Godart B, Lenoir B, Leveille A, Neron P and Tonnoir B (1989) Viaduc dAmiens. Conception
et realisation de louvrage. Suivi experimental sur site, pendant la construction. Annales
ITBTP, MarchApril.
Halpin DW and Riggs LS (1992) Planning and Analysis of Construction Operations. Wiley, New
York, USA.
Hetenyi H (1971) Beams on Elastic Foundation, Theory with Application in the Fields of Civil
and Mechanical Engineering. University of Michigan, Ann Arbor, MI, USA.
Leonhardt FC (1979) C.A. e C.A.P. Calcolo di Progetto e Tecniche Costruttive. Vol. 4 I Ponti,
Dimensionamento, Tipologia, Costruzione. Edizioni di Scienza e Tecnica, Milan, Italy.
Llombart JA and Revoltos J (2000) Petra Tou Romiou Viaduct, Cyprus. Structural Engineering
International (10)4: 233234.
Mapelli M, Mola F and Pisani A (2006) Time-dependent analysis of launched bridges.
Structural Engineering Mechanics 24(6): 741764.
Marzouk M, El-Dein HZ and El-Said M (2007) Application of computer simulation to
construction of incremental launching bridges. Journal of Civil Engineering and Management
XIII(1): 2736.
Marzouk M, Said H and El-Said M (2011) Framework for multi-objective optimization of
launching girder bridges. Journal of Construction Engineering and Management 135(8): 791
800.
Ontario (2006) Incrementally Launched Post-tensioned Concrete Bridge Design, May.
Pritsker AAB, OReilly JJ and Laval DK (1997) Simulation with Visual SLAM and AweSim.
Wiley, New York, USA.
Ramakrishna A and Sankaralingam C (2000) Panval Nadhi Viaduct, India. Structural
Engineering International (7)3: 168170.
Rosignoli M (1981) Ponti a cassone: la distorsione della sezione trasversale secondo la analogia
della trave su mezzo elastico. University of Ancona, Ancona, Italy.
Rosignoli M (1996) Sul dimensionamento degli impalcati da ponte in c.a.p. realizzati per varo
frontale progressivo. LIndustria Italiana del Cemento, December.
Rosignoli M (1997a) Inuences of the incremental launching construction method on the sizing
of prestressed concrete bridge decks. Proceedings of the ICE Structures and Buildings
122(3): 316325.
Rosignoli M (1997b) Tolleranze di posizionamento degli appoggi di varo nei ponti spinti in
c.a.p. LIndustria Italiana del Cemento, September.
Rosignoli M (1998a) Serio River Bridge, creep and incremental launching. Proceedings of the
ICE Structures and Buildings 128(1): 111.
Rosignoli M (1998b) I tre ponti in c.a. precompresso del nodo di Via Palizzi. Le Strade, June.
Rosignoli M (1998c) Site restrictions challenge bridge design. Concrete International 20(8).
Rosignoli M (1998d) Misplacement of launching bearings in PC launched bridges. Journal of
Bridge Engineering 3(4): 170176.
Rosignoli M (1998e) Launched Bridges. ASCE Press, Reston, VA, USA.
Rosignoli M (1999a) Prestressing schemes for incrementally launched bridges. Journal of Bridge
Engineering 4(2): 107115.
Rosignoli M (1999b) Presizing of prestressed concrete launched bridges. ACI Structural Journal
96(5): 705710.
Rosignoli M (2000) Thrust and guide devices for launched bridges. Journal of Bridge
Engineering 5(1): 7583.

223

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


Bridge Launching

Rosignoli M (2001) Deck segmentation and yard organization for launched bridges. Concrete
International 23(2).
Rosignoli M (2008) Bridge launching and shifting on the Tanaro river, Alessandria, Italy.
Proceedings of the 1st ASBI International Symposium on Future Technology for Concrete
Segmental Bridges, San Francisco, CA, USA.
Rosignoli M (2010) Incremental launching of prestressed concrete urban bridges. Proceedings
of the 8th International Conference on Short and Medium Span Bridges, Niagara Falls,
Canada.
Rosignoli M (2013) Bridge Construction Equipment. ICE Publishing, London, UK.
Rosignoli C and Rosignoli M (2007) Launch and shift of the Tiziano Bridge. Concrete
International 29(10).
Rosignoli C and Rosignoli M (2008) Incremental launching construction of urban bridges.
Proceedings of the 2008 IABSE Congress, Chicago, IL, USA.
Seifried G and Wittfoht H (1979) Die Brucke uber den Shatt-al Arab in Basrah (Iraq). Beton
und Stahlbetonbau 74: 7785.
Shannon RE (1992) Introduction to simulation. Proceedings of the 1992 Winter Simulation
Conference, Arlington, VA, USA, pp. 6573.
Wright RN, Abdel Samad SR and Robinson AR (1968) BEF analogy for analysis of box
girders. ASCE Journal of the Structural Division 94(ST7): 17191743.

224

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

Potrebbero piacerti anche