You are on page 1of 5

Monolithic Launch

of the Reggiolo
Overpass
An example of an innovative construction method
By Marco Rosignoli

ven though the Reggiolo Overpass in Reggiolo, Italy, is


a small bridge, several factors complicated its design
and construction. The bridge, completed in 2003, spans
the Verona-Mantua Railway that had to remain fully
operational during construction. This made the formwork
and shoring that would have been necessary for the
originally designed cast-in-place bridge difficult to locate
and construct. Its location over a working railway also led
to the owners desire for minimal long-term maintenance,
resulting in the selection of a monolithic, fully prestressed
structure instead of a steel or precast/prestressed
girder system.
Settling soils at the site also affected the design. The
structure was simply-supported so that higher-thanexpected foundation settlements could be recovered by

shimming at the abutments, and the structure was


lightened by using high-density polystyrene blocks to fill
voids between the top and bottom slabs. These blocks
also eliminated empty cells that require inspection and
can gather water, further reducing maintenance. The
bridge also had a complex, trapezoidal plan geometry as
a result of the entry and exit ramps that converged on
the overpass.
The demanding requirements were met by casting
the superstructure on one side of the span, and then
monolithically launching it over the railway into its final
position. Similar construction techniques have been
adopted for Italian railway overpasses,1 but in this case,
the small dimensions and complex geometry suggested
monolithic, rather than incremental, launching.

Deck design

Fig. 1: Plan view of the internal stiffening systems in the deck plate
(All dimensions in meters; 1 m = 3.28 ft)

As shown in Fig. 1, the superstructure of the overpass


is a multi-cellular, prestressed concrete plate spanning
26 m (85.3 ft). The bottom surface is horizontal in the
transverse direction and slightly inclined longitudinally,
but the top surface is inclined in both directions to
shed water, so that the total thickness of the structure
varies from 1.28 to 1.35 m (4.2 to 4.4 ft), with a mean
span-to-depth ratio of about 20. A varying-depth paving
layer would have been excessively heavy on such
a wide deck.
In the transverse direction, the plate is stiffened by
three 0.4 m (15.7 in.) thick internal diaphragms spaced
at 6.5 m (21.3 ft) and two 1.0 m (3.3 ft) thick end beams
at the abutments. A section at the midspan diaphragm
is shown in Fig. 2. Because the launch supports were
aligned with the three internal longitudinal webs and
the side cantilevers ranged from 3.24 to 6.30 m (10.6 to
Concrete international

/ February 2007

61

20.7 ft), a high level of prestress in the transverse direction


was required during launching. In addition, the transverse
negative moment was increased by the concentrated load
of the edge beam.
The prestressing level in the superstructure was
quite high because of the railway authoritys request for
a fully prestressed section in both the longitudinal and
transverse directions under full design loads (dead
and superimposed dead loads, live loads inclusive of
dynamic amplification, and thermal gradients) and also
during launching. In spite of the short construction
duration, consideration of time-dependent launch
stresses2 was required.
The prestressing included: a) longitudinal launch
tendons in the internal webs and edge beams; b)
longitudinal tendons in the bottom slab, tensioned
after launch completion; c) transverse tendons in the
end beams and internal diaphragms, all tensioned
before launching; and d) transverse tendons in the
deck slab, also tensioned before launching.

Deck construction

After casting the foundation piles and the two


abutments, the casting yard was built on the embankment
leading to the launch abutment. It consisted of three
parallel 1.8 m (5.9 ft) deep foundation beams spaced
5.5 m (18.0 ft) apart to match the spacing of the three
internal webs (Fig. 3).
The foundation beams were anchored to the launch
abutment to minimize settlement and transfer the launch
forces to the abutment. Figure 4 shows the deck after it
was lifted from the foundation beams for prelaunch
weighing and shows the wide cantilever at the rear deck
section. Before launching, the deck was painted and
equipped with safety fences and guard rails.
The three prestressed concrete launching noses
at the front of the deck (Fig. 5) were constructed to
control negative moments at the front of the bridge by
shortening the cantilevered span that the deck would
have to carry before reaching another support. They
also limited the safety factor for deck overturning about
the launch abutment to a minimum
of 1.5. The noses were 3 m (9.8 ft)
long, 0.60 m (2 ft) thick, and 1.15 m
(3.8 ft) deep and aligned with the
three internal webs. As shown in
Fig. 5, each nose was anchored to
the corresponding deck web with
eight 36 mm (1.4 in.) prestressing
bars. These overlapped the ends
of the four launch tendons (Fig. 1).
Shear keys on the contact surface
between the deck and noses
were match-cast to transfer vertical
Fig. 2: Cross-section of the deck slab at the midspan diaphragm (All dimensions in
meters; 1 m = 3.28 ft)
forces from the nose to the deck.
Steel braces joined the three
noses together.
Short prestressed concrete noses
such as these, that are detached and
demolished after launch completion,
are often less expensive than
conventional steel girders, especially
when the negative moments are too
high to allow full-span launching of
the prestressed deck.3 As shown in
Fig. 6, the high negative moments
were alleviated by cutting the launch
span in half using a temporary, threetower steel pier placed near the
railway. Each tower was vertically
supported by a foundation pile, and
the tops of the towers were anchored
to the launch abutment with rigid
stays that resisted the horizontal
Fig. 3: Support points of the deck plate during launching (All dimensions in meters;
1 m = 3.28 ft)
forces from sliding friction.

62

February 2007

/ Concrete international

Because of its varying width,


the deck alignment could not be
maintained with conventional lateral
guides during launching.4 Instead,
guide pins at the rear of the deck and
at the launch abutment were used.
These pins guided the bridge through
a steel-lined recess in the underside
of the deck (Fig. 2). At the temporary
pier, the central guide was used to
laterally stabilize the pier through
the deck.
After tensioning the launch
prestressing tendons (transverse
tendons in the end beams and the
internal diaphragms, transverse
slab tendons, and longitudinal launch
tendons), the deck plate was lifted
off the foundation beams using three
rows of hydraulic sliding supports
aligned with the foundation beams
and the internal deck webs (Fig. 3).
Because the launching tendons were
tensioned while the structure was
resting on the foundation beams,
a portion of the prestressing force
may have been lost to friction
between the structure and the
foundation beams. This required
a final check of the prestressing
force after lifting the structure. The
strands were retensioned if found to
be at less than 98% of the design
value. The deck was also weighed to
confirm its theoretical weight, and
measured deflections were compared
to theoretical deflections for an
indirect evaluation of the modulus of
elasticity of the concrete.5
With the bridge supported on 2 or
3 rows of hydraulic sliding supports,
hydraulic pistons acting between
the deck and the foundation beams
(Fig. 7) pushed the bridge forward
out over the span. After launching
the desired distance, the support
points for the bridge were changed,
and the pistons were repositioned.
This process was repeated in
several stages, as illustrated in
Fig. 8, until the bridge reached its
final position.
After releasing and dismantling
the launch noses, the bottom slab

Fig. 4: Deck weighing before launching through transducer-controlled hoisting. The


four blockouts in the foundation beam house the hydraulic jacks at the support
points shown in Fig. 3

Fig. 5: The prestressed concrete launch noses were connected with steel bracing
and were used to control negative moments while the bridge was launched. They
were removed after the bridge was in its final location

Fig. 6: Temporary steel towers were used to cut the span of the bridge in half
during the launch. The steel braces connected the top of the tower to the launch
abutment and resisted the lateral forces at the support points

Concrete international

/ February 2007

63

Fig. 7: Paired launch pistons and central anti-drift guide at the rear deck end

tendons were tensioned and grouted, the supports at the


temporary towers were removed, and the bridge was
placed on permanent laminated-rubber bearings at the
two abutments.

Additional considerations and comments

The temporary stresses in the deck plate during


the monolithic launch were analyzed at the beginning
and end of each launch stage and at intermediate
stages where necessary. A three-dimensional grillage
model was used for this analysis because a twodimensional beam model would have lead to excessively
approximate results.6 The analyses also accounted for
settlement of the temporary pier and the prescribed
tolerances in the support reactions in both the
longitudinal and the transverse direction. Comparison
of the results of launching and service load analyses
showed that positive moments in the longitudinal webs
were governed by service conditions, while launching
governed negative moments (these negative moments
would have been totally absent in the case of conventional
deck construction on shoring). Although these temporary
stresses increased the cost of prestressing the deck, the
erection, dismantling, and materials for scaffold-type
shoring on piles and the higher embankments required
for working clearance above the railways would have
generated about 12% higher costs.

64

February 2007

/ Concrete international

The launch equipment was particularly inexpensive.


Because of the bridges modest weightabout 8900 kN
(1000 ton)a pair of small long-stroke launch pistons
(Fig. 7) generated adequate operational speed at minimal
cost.4 The use of hydraulic launch supports in the casting
yard instead of conventional continuous low-friction
extraction rails7 generated substantial savings in this
bridge because three 27.6 m (90.6 ft) long steel rails
would have been particularly expensive. In addition, deck
lifting permitted real-time monitoring of the support
reactions during launching, as well as compensating for
the settlements of the support embankment.
To further save manpower, new types of sliding supports
were tested at the temporary pier, where Teflon plates
were placed directly onto the hoisting jacks and long
stainless steel sheets were inserted between the hoisting
jacks and the bridge for long-stroke continuous sliding
(Fig. 6). The result was encouraging, although loss of the
compensating effect of the rubber sheets of conventional
low-friction bearing pads is acceptable only when hydraulic
launch supports are used, and the nonpolished,
nontensioned surface of the stainless steel sheets
unavoidably increased sliding friction.
Accurate load testing of the deck plate after launch
completion confirmed the absence of any nonlinearity as
a result of the launch process, and the prescribed
absence of longitudinal and transverse tensile stresses

completely avoided by casting the superstructure on stiff


foundation beams with inexpensive forms,9 thus improving
the quality of the structure.

Acknowledgments
The Reggiolo Overpass was built by the firm Locatelli, SpA, for
the owner Autostrada del Brennero, represented by G. Andreani.
The general designer of the motorway bypass was Sembenelli
Consulting. The author designed the prestressed deck plate for both
launch and service load stages, performed independent design
checking of the launch equipment for the owner, and supervised the
launch process and final load testing.

References
1. Rosignoli, M., Site Restrictions Challenge Bridge Design,
Concrete International, V. 20, No. 8, Aug. 1998, pp. 40-43.
2. Rosignoli, M., Creep Effects During Launch of the Serio River
Bridge, Concrete International, V. 22, No. 3, Mar. 2000, pp. 53-58.
3. Rosignoli, M., Nose-Deck Interaction in Launched Prestressed
Concrete Bridges, Journal of Bridge Engineering, V. 3, No. 1, Feb.
1998, pp. 21-27.
4. Rosignoli, M., Thrust and Guide Devices for Launched
Bridges, Journal of Bridge Engineering, V. 5, No. 1, Feb. 2000, pp. 75-83.
5. Rosignoli, M., Launched Bridges, ASCE Press, Reston, VA, 1998,
363 pp.
6. Rosignoli, M., Reduced-Transfer-Matrix Method for Analysis of
Launched Bridges, ACI Structural Journal, V. 94, No. 4, July-Aug.
1999, pp. 603-608.
7. Rosignoli, M., Bridge Launching, Thomas Telford, Ltd., London,
2002, 342 pp.
8. Rosignoli, M., Incremental Bridge Launching, Concrete
International, V. 19, No. 2, Feb. 1997, pp. 36-40.
9. Rosignoli, M., Deck Segmentation and Yard Organization for
Launched Bridges, Concrete International, V. 23, No. 2, Feb. 2001,
pp. 64-73.
Selected for reader interest by the editors.

Fig. 8: Launch sequence for the Reggiolo Overpass

during launching was monitored in real time through


hydraulic control of the support reactions and comparison
with their theoretical values.

Safe and reliable

This application of the monolithic-launch construction


method demonstrated its reliability and safety.8 Adoption
of this method allowed a delicate and complex bridge to
be built with extreme safety for the workers, in a short
timeframe, and without any restriction on the railway
traffic below. The adverse effects of soil settlement were

Marco Rosignoli recently joined HNTB Corp.


as Chief Bridge Engineer in the Seattle, WA,
office after working 8 years as a Site
Manager for bridge projects, 7 years as the
Technical Director for a major Italian bridge
contractor, and 9 years as a freelance
bridge consultant. He has authored two
books and several publications on the
incremental-launching construction of
bridges and also provides independent design checking of
launching gantries and movable formwork systems.
Concrete international

/ February 2007

65

You might also like