Sie sind auf Seite 1von 9

DC Iles page 1

DESIGN ILLUSTRATION COMPOSITE HIGHWAY BRIDGES


DC Iles, The Steel Construction Institute, Ascot, UK

Abstract
Design of a composite highway bridge to the Eurocodes will require reference to at least
14 separate Parts of the Eurocodes, each with its appropriate National Annex. To illustrate
many of the aspects of applying the necessary documents to the design of typical multi-girder
and ladder deck bridge configurations, SCI has published a book with two worked examples.
This paper presents an overview of some of the design aspects revealed in preparation of
those examples.

Introduction
For the last 20 years, SCI has provided guidance to the designers of composite highway
bridges. General guidance on best practice, based on the views of experienced senior
designers, has been accompanied by guidance on the use of design standards, notably, in the
past, on the use of BS 5400. The guidance has been illustrated by worked examples,
presenting detailed calculations, with references to the relevant clauses of the standards. That
guidance has now been updated with two new publications for design in accordance with the
Eurocodes, one that offers general guidance (SCI publication P356)[1] and one that presents
two worked examples (SCI publication P357)[2]. In P357, one example is a multi-girder bridge
and the other is a ladder deck bridge. The preparation of the examples has revealed many of
the aspects where design practice differs, to a greater or lesser degree, from that in accordance
with BS 5400. This paper presents an overview of those aspects.

The Examples
In SCI publication P357, Example 1 is a two-span multi-girder deck bridge with integral
abutments. The arrangement of the bridge is shown in Figure 1. Example 2 is a three-span
ladder deck bridge, with a curved soffit to the main girders in all three spans. The
arrangement of the bridge is shown in Figure 2.
DC Iles page 2
28000 28000

500 2000 1000 7300 1000 2000 500


marginal marginal
strip strip

1100

3700 3700 3700

Figure 1. General arrangement of Example 1

24500 42000 24500

500 2500 1000 7300 1000 2500 500


marginal marginal
strip strip

varies
1200 to 2200

11700

Figure 2. General arrangement of Example 2

Preamble to Design Verification


Documentation
The division of the Eurocodes into numerous parts, each representing a separate subject,
results in the requirement to refer to at least 14 Parts, possibly as many as 20 Parts, for the
design of a composite highway bridge. In addition to EN 1990, for the basis of design,
reference is needed at least to Eurocode 1 (Parts 1-1, 1-5, 1-6 and 2), Eurocode 2 (Parts -1-
and 2), Eurocode 3 (Parts 1-1, 1-5, 1-8, 1-9, 1-10 and 2) and Eurocode 4 (Part 2). Foundation
design will require reference to Eurocode 7. Other Parts will be needed for more complex
structures, such as long-span cable-stayed bridges.
DC Iles page 3

For structures in the UK, the UK National Annexes must also be consulted. Some of these
National Annexes refer to BSIs published documents (PDs), which are one form of
non-contradictory complementary information (NCCI). NCCI has no special status
according to the Eurocodes, its merely what it says it is: a text book may well be considered
as NCCI; industry is free to produce documents that may be so termed NCCI (as long as they
do not contradict).

Design basis
The categorisation of design situations and combinations of actions to be considered is well
set out in EN 1990 and is readily applied for ordinary highway bridges. The terminology used
is clear and the use of Ed and Rd as subscripts for design effects and design resistances
contributes to clarity in many aspects of the verification procedure. One minor oversight in
the UK NA to BS EN 1990[3] is the omission of values for partial factors during transient
situations (notably construction), although the use of factors for persistent situations should be
conservative.

Material properties
Properties of steel and concrete material properties are clearly defined in EN 1993-1-1 and
EN 1992-1-1 respectively. For concrete, the determination of long-term modulus of elasticity
and shrinkage depends on the project-specific parameters of relative humidity and age at
loading. No doubt, a common practice will evolve on normal assumptions for these
parameters at opening to traffic and later in the working life. In the examples, where
shrinkage effects are favourable they are neglected, rather than using the lesser values at the
time of opening to traffic.

The use of different coefficients of expansion for uniform temperature change and for
temperature gradient will be puzzling to many but is easily dealt with.

Design Effects
Actions
Values for the density of steel and concrete are given in EN 1991-1-1. It is reasonable to use
the lower end of the range for steel self-weight (77 kN/m3); the suggested addition of 1 kN/m3
for the self weight of reinforcement may be rather low for typical reinforcement in deck slabs
and the designer should adopt a suitable value. For surfacing, the allowance for an additional
thickness of 55% of nominal value given by the UK NA to BS EN 1991-1-1, Table NA.1 is
more onerous than the values given by BS 5400.

Strictly, the traffic load UDL (in Load Model 1, see EN 1991-2:2003, 4.3.2) should only be
applied to adverse areas of influence surface and these do not necessarily align with
boundaries between traffic lanes. A more practical approach of applying the UDL over full
lane widths, rather than part widths, will make very little difference to design effects. The use
of a fixed width of notional lane (rather than dividing the available width into an integral
number of lanes) and remainder widths that are also loaded may seem a little odd but should
cause little difficulty in practice. Thanks to the UK NA, the UDL has the same intensity over
the full adverse area. Since the UDL does not vary with loaded length, it is simpler to apply
than BS 5400 HA loading.
DC Iles page 4

The single vehicle fatigue load model defined in EN 1991-2:2003, 4.6.4 is used in the
simplified assessment of fatigue in EN 1993-2:2006, 9.2.2. Although EN 1993-2:2003, 9.5.2
implies that a spectrum of lorry weights is needed for the simple assessment, no appropriate
spectrum is offered in EN 1991-2. The UK NA to BS EN 1993-2 avoids the need for defining
a suitable spectrum by giving an average weight for UK traffic (the average is independent
of the type of road). It is not entirely clear whether the fatigue load model should be applied in
the same notional lanes as for Load Model 1 or in the marked lanes (as BS 5400 required).
The reference in the UK NA to numbers of vehicles in slow and fast lanes (which are not
defined) would indicate the use of the marked lanes. The use of notional lanes would be
conservative.

Global Analysis
According to both Eurocodes 3 and 4 elastic global analysis is to be used for bridge structures,
although the UK NA to BS EN 1993-2:2006, NA.2.16 does allow particular projects to
specify where plastic analysis would be acceptable (such as for accidental situations). No
guidance is available on where plastic analysis might be appropriate.

In the elastic global models, both EN 1993-2 and EN 1994-2 require the use of effective
widths of wide flanges, allowing for shear lag. Simplified approaches are given. However, if
FE models are used, with shell elements for the deck slab, shear lag will be automatically
taken into account and no allowance should be made in the model. However, this means that
moments on notional composite beams extracted from such a 3D model should be determined
from the stresses in the gross flange width, for verification against the resistance of a cross
section that includes only the relevant effective width.

Traditionally, first-order (small deflection) analysis models have been used and buckling
resistance has been determined by reference to empirical rules for effective length. With the
increasing power of modern software, elastic buckling analysis of a 3D model is now possible
and this offers advantages in some situations. However, the interpretation of output from such
software does require experience to ensure that appropriate buckling modes have been
identified.

The models used in the Examples were 3D FE models (first-order) using shell elements for
the deck slab and beam elements for the flanges, stiffeners and bracing members. An
illustration of the ladder deck model, for a construction stage with only part of the deck slab,
is shown in Figure 3.
DC Iles page 5

Figure 3. Global analysis model for ladder deck bridge (construction stage)

Design Verification of Beams


Resistance of cross-sections
Although Eurocodes 3 and 4 consider 4 classes of cross section, compared to the compact
and non-compact designations in BS 5400, there is no significant change, since Classes 1
and 2 are treated in the same way for bridges and correspond to compact sections. Class 4 is
equivalent to a non-compact section with a web that is not fully effective (although
Eurocode 3 takes a slightly less conservative view than BS 5400-3 of the loss of
effectiveness).

When considering the design resistance of composite sections, care must be taken with the
value of the design strength of concrete the Eurocode 2 definition includes the parameter cc
which, according to the UK NA has a value of 0.85; the design value is thus based on 85% of
the cylinder strength whereas the Eurocode 4 definition bases it on 100% of the cylinder
strength (but only mobilizes 85% of that strength on the compression side of the plastic
neutral axis).

Shear resistance (of a beam web) is considered to be a property of the cross-section but it does
depend on the stiffening along the beam. Where shear buckling resistance is mobilized, the
partial factor M1 must be applied, in contrast to the M0 factor applied for design resistance of
the cross section in bending. (The values of the factors are M0 = 1.0 and M1 = 1.1, according
to the UK NA to BS EN 1993-2: 2005.)

Bending/shear interaction is only considered in relation to the bending resistance of the


cross-section, not in relation to the buckling resistance of the member (as in BS 5400).
Generally the limiting envelope is similar to that in BS 5400 except that is curved, rather than
polygonal (see Figure 4).
DC Iles page 6

3500
V b,Rd
3000
M el,Rd

Shear (kN) 2500 V bw,Rd

2000

1500

1000

500
M f,Rd M pl,Rd
0
0 5000 10000 15000 20000
Moment (kNm)

Key: Maximum shear with coexisting moment


Maximum moment with coexisting shear

Figure 4. Limiting envelope for moment-shear interaction (Example 1)

One aspect not mentioned in the Eurocodes is the resistance of non-parallel sections, such as
with a haunched profile or a curved soffit. In such sections some of the shear is carried by a
component of the force in the inclined flange. It seems reasonable to continue the practice of
reducing the design value of shear on a cross section (VEd in Eurocode terminology) by the
vertical component of force in the compression flange. Slenderness of the web in shear should
be based on the deeper end of a tapered web panel.

Buckling resistance of beams


For composite highway bridges, the two regions where buckling resistance needs to be
verified are midspan regions during construction (at the wet concrete stage) and adjacent to
intermediate supports (at the in-service stage) the hogging moment regions. The first of
these involves true lateral torsional buckling (albeit with flexible intermediate torsional
restraints provided by bracing or cross girders). The second is actually lateral distortional
buckling, though EN 1994-2:2005, 6.4.2 says that the resistance for this mode of buckling
may be evaluated using the reduction factor for LTB given in EN 1993-1-1 for the appropriate
non-dimensional slenderness and buckling curve.

At the wet concrete stage, beams in multi-girder bridges are usually paired together and the
LTB buckling mode involves the pair of beams twisting and displacing as a unit. Similarly,
the main girders in a ladder deck are paired together by the cross girders. The mode and
critical buckling load are influenced by the stiffness and spacing of bracing, and by the
variation in bending moments and section properties across the span. The most accurate
means to determine the elastic critical load, and thence the non-dimensional slenderness and
reduction factor for LTB that is to be applied to the resistance of the cross section, is by a 3D
elastic buckling analysis. However, such analysis is not always available and even when it is,
interpretation of output requires care and experience. One alternative is to adopt the empirical
rules from BS 5400-3, which determine first an effective length for buckling and from that
the non-dimensional slenderness. The rules are available in SCI publication P356[1] and in
PD 6695-2[4]. The Examples in P357 used the empirical rules and Figure 5 shows the
DC Iles page 7

deflected form of the ladder deck model when determining the torsional stiffness of the beams
in the central span, as the first step to determining slenderness.

Figure 5. Displaced shape of bare steel ladder deck due to unit moments about the
longitudinal axes applied to both main beams over the central span

For hogging moment regions adjacent to supports, the elastic critical load for buckling of the
bottom flange can also be determined using a 3D elastic buckling analysis and the beam
verified using the general method of EN 1993-2:2006, 6.3.4.1. However, EN 1993-2:2006,
6.3.4.2 offers a very useful simplified strut model method. The flange plus part of the web is
treated as a T-shaped strut and its lateral slenderness evaluated, making allowance for
non-uniform force (along its length) and for flexible lateral restraint from the web. This is in
effect a type of beam on elastic foundation model, as employed in the design of U-frames in
BS 5400-3. In Example 1, the cross bracing (at about the quarter point of the span) was
sufficiently stiff to be an effective lateral restraint and the slenderness was determined for the
T-shaped flange plus part of the web, as a strut over the distance from support to bracing. In
the ladder deck model the U-frame created by the first cross girder from the support and the
web stiffeners to which it is attached was not sufficiently stiff to be an effective restraint to a
strut over the distance from the support. The stiffness requirement at the second frame (which
is less onerous because the strut is twice as long) was satisfied by the U-frame at that position
and the slenderness was calculated for this longer length. The bottom flange in this case is
curved and the depth of the beam decreases over the buckling length; this does not affect the
determination of slenderness (the strut buckles laterally) but does require careful
consideration of the values of moment and resistance at appropriate cross sections (both vary
along the beam).

Restraint of deck slabs in ladder deck bridges


In the span regions of ladder deck bridges, the deck slab is effectively a wide plate in
compression that is restrained against buckling, out its plane, by the cross girders. The cross
girders thus need to be stiff and strong enough to provide that restraint; there are no explicit
rules in the Eurocodes for the requirement but it is possible to give guidance that is
compatible with other rules.
DC Iles page 8

For the usual spacing of cross girders (about 3.5 m centres) the slab itself is sufficiently
slender that second order effects in the slab would need to be allowed for, in accordance with
EN 1992-1-1, although that verification is not given in P357.

Longitudinal shear connection


The strength of shear stud connectors to the new European Standard (EN ISO 13918) is about
10% less than the old BS 5400 studs but the design value of their shear resistance at ULS is
comparable. As SLS, the limiting value of shear force per connector is 75% of the ULS shear
resistance but since the design effects are typically no more than about 75% of the ULS
values, SLS does not often govern the main exception would be at the ends of the bridge,
where primary forces due to temperature difference have to be transferred over a short length
of girder).

In a 3D analysis, the proper 3D behaviour of the structure as a whole is modelled and thus,
with loading that is not uniform across the width of the bridge, the notional composite beams
(steel girder plus part of the deck slab) carry both bending and axial forces. The axial forces
vary along the span and this variation may add to the longitudinal shear between the girder
and the slab; in the Examples, the variation of axial force was taken into account, although it
was modest in magnitude.

Design of Connections
Bolted connections
Apart from a change of terminology (from HSFG bolts acting in friction to slip-resistance
shear connections) the rules in Eurocode 3 are similar to those in BS 5400. Bolts for
preloading are now supplied to EN 14399, usually in size M24 or M30. Ordinary (hexagon
head) bolts are available in Grades 8.8 and 10.9; type HRC (the generic designation for TCB))
is available in Grade 10.9.

Example 1 includes the design of a typical bolted splice. The flange and web cover plate
connections were designed to resist the maximum forces at the splice position. The axial force
in the top flange was found to be greatest, and compressive, at the wet concrete stage are
consequently was magnified to allow for second order effects due to lateral buckling of the
flange.

Example 2 includes the design of the lapped connection at the ends of cross girders. The
connection detail and the model of a width of concrete slab plus a steel web are shown in
Figure 6. The most onerous requirements on this connection are adjacent to the intermediate
support, where U-frame forces are developed due to loading on the cross girder and due to
restraint of the compression flange. Determination of the design force on the extreme bolt
depends on judgement of effective width of slab and thus the centre of rotation for the bolt
group; in the absence of appropriate rules for effective width, a pragmatic judgement was
made and the connection was designed against slip at ULS.
DC Iles page 9

250 mm

(notch ignored)

725 mm

Figure 6. Main girder / cross girder connection in ladder deck bridge

Welded connections
The Examples verify the adequacy of the web/flange welds and the welds at the bottom of the
bearing stiffeners, for static and fatigue loading. Weld sizes are expressed as throat thickness,
which is consistent with the detailed rules in EN 1993-1-8, rather than leg length. When
specifying weld size on drawings or other contract documents, it must be made clear when
throat size is give, rather than the traditional UK practice of specifying leg length.

Conclusion
Preparation of the two worked examples revealed no major differences in design procedures
(from those to BS 5400). Terminology is slightly different (and rather more precise) and
procedures are sometimes a little more exacting. One omission from the Eurocode rules - the
absence of simple expressions to calculate elastic critical moment, rather than resorting to a
buckling analysis - is catered for either by use of its simplified strut model or by use of
NCCI adapted from rules in BS 5400. There is no indication that the Eurocode rules require a
heavier structure than do the BS 5400 rules.

References
[1] Iles, D.C. (2010). Composite highway bridge design, Publication P356, The Steel
Construction Institute.
[2] Iles, D.C. (2010). Composite highway bridge design: Worked examples, Publication
P357, The Steel Construction Institute.
[3] NA to BS EN 1990:2002+A1:2005, UK National Annex for Eurocode Basis of
structural design (Amd A1, 2009), BSI
[4] PD 6695-2:2008, Recommendations for the design of bridges to BS EN 1993, BSI

Das könnte Ihnen auch gefallen