Vous êtes sur la page 1sur 31

CT5125: Steel bridges – file <design-box-girder>

20. Box Girder Bridge

20.1 Introduction
IThe nomenclature of the structural elements in a steel box girder is given in figure 1 which shows, as an
example, a single cell box girder with a composite concrete deck.
Until 1940 the structural possibilities for box girders were
limited; structures had to be assembled from rolled sections,
plates and riveted connections.

Fig. 1.
Box girder bridge with composite
concrete deck: nomenclaturen

Notwithstanding these limitations, the first steel box girder, the


Conway bridge with a single span of 125.5 m and the Britannia
Bridge (1850) with main spans 152 m, figure 2, served as a model
of what could be achieved with innovative design.

Fig. 2. Left: The Conway bridge (tubular bridge).


Middle and right: The Britannia bridge.

The basic concept of using hollow sections was only occasionally repeated with riveted construction.
• The Britannia Bridge was duplicated only once; in America. A box girder is more material consuming
than a truss girder, and material was much more expensive than labor in those days.
• The tubular members of the Firth of Forth Bridge (1890) were a second exception.
• The railway bridge across the Oude Maas, Dordrecht, The Netherlands, has tubular riveted
diagonals. Here a corrosion problem arose. Riveted boxes are not completely watertight.
Humid airs sucked in, condensed and water collected at the bottom.

With the development of electric welding and precision flame


cutting, the structural possibilities increased enormously. It is
now possible to design large welded units in a more
economical way, e.g. box girders, using the techniques similar
to those of shipbuilding.
A box girder consists of:
• a concrete deck or an orthotropic steel deck as the
top flange, and sometimes a combination of the two
• a stiffened plate or a bracing as a bottom flange
• web vertical or inclcined
• stiff diaphragms or bracing’s at the supports and
lighter cross bracing’s between the supports at
distances of about 2,5 times the construction depth,
Fig. 3. Section of a steel box girder bridge.
Nowadays, the basic cross-section can be found in many bridges.

20. Design of a steel box girder 301


Dr. A. Romeijn
CT5125: Steel bridges – file <design-box-girder>

The great torsion rigidity makes a box girder a particularly appropriate solution where the bridge is curved in the
horizontal plane, figure 4. Many bridges over European highways may serve as examples. Launching as an
erection method is then still possible as long as the curvature is constant.

Fig. 4.
Left: Plate girder bridge.
Right: Box girder bridge.

In wide cross-sections the box is sometimes subdivided


into cells, figure 5.

Fig. 5. Types of cross section (concrete – composite concrete / steel – steel).

Nowadays, composite steel-concrete box girder bridges are commonly used in (curved) bridges. They generally
consist of one or more U-girders attached to a concrete deck through shear connectors. Diaphragms connect
individual steel U-girders periodically along the length to ensure that the bridge system behaves as a unit. The
cross section of a steel box is flexible (i.e. can distort) in the cross-wise direction and must be stiffened with
cross-frames that are installed in between the diaphragms to prevent distortion. Web and bottom plates stiffeners
are required to improve stability of the relatively thin steel plates that make up the steel box. During construction,
overall use of top bracing members enhances stability and torsional rigidity of the girder. These bracing members
become unimportant once the concrete deck hardens, but is left in place anyway.

20. Design of a steel box girder 302


Dr. A. Romeijn
CT5125: Steel bridges – file <design-box-girder>

Fig. 6.
Example composite steel-concrete bridge using diaphragms connecting
individual steel U-girders.

Fig. 7. Left: Transportation of box section with stiffened flange.


Right: Welded connection on site.

Fig. 8.
Hybrid box girder: concrete at support region and steel at
mid-span.

20. Design of a steel box girder 303


Dr. A. Romeijn
CT5125: Steel bridges – file <design-box-girder>

Fig. 9. Hemboog: railway viaduct Zaandam – Amsterdam, total length


2.2 km.

Fig. 10. Silent bridge technology (Holland Railconsult).

20.2 General Design Principles

Span
Box girders are suitable for longer spans than I-girders and allow larger span to depth ratios. The limits for
competitiveness may vary due to local market conditions.
Steel or steel-concrete composite box girders are usually more expensive than plate girders because they require
more fabrication time. They have, however, several advantages over plate girders which make their use
attractive:
• very high torsional rigidity: In closed box girders, torque is resisted mainly by Saint Venant shear
stresses because the Saint Venant torsional stiffness is normally much greater than the torsional
warping stiffness. For highly curved spans, this stiffness of the box girders is virtually essential during
their construction, as well as under service loads. All steel box girders provide torsional stiffness during

20. Design of a steel box girder 304


Dr. A. Romeijn
CT5125: Steel bridges – file <design-box-girder>

their erection. Composite box girders only achieve their torsional rigidity after concreting. During
erection and concreting they may require expensive temporary bracing, which can also interfere with
the execution of the concrete slab.
• very wide flanges allow large span-to-depth ratios
• a neater appearance, since the stiffening can remain invisible in the box
• very good aerodynamic shape, which is equally important for large suspension or cable-stayed bridges
as is the torsional stiffness
• a very good adaptability to the most difficult conditions. Box girders are able to cross greater torsional
spans than flexural spans using piers with a single bearing as shown in Figure 4.
• low noise emission and reduction on costs for corrosion protection
• easy handling

Table 1 shows an indication of economic span limits for road bridges.

Table 1. Span range for box girder bridges.


Composite concrete deck (m) Orthotropic deck (m)
Simple span 20 - 100 70 - 120
Interior span of continuous 30 - 140 100 - 250
girder

The longest span so far is 300 m achieved in 1974 by the Costa e Silva bridge in Rio de Janeiro, figure 11. It is
always probable that the longest span existing has passed the limit of best economy.

Fig. 11. Rio-Niteroi bridge.

Span-to-depth ratio
The span-to-depth ratio will normally be around 20 to 25 for simple girders and around 25 to 35 for continuous
girders. It is possible to reduce the depth, if necessary, without violating deflection limitations, at the expense of
additional steel. The above ratios are valid for road bridges. For rail bridges the ratios should be smaller, say 15
and 20. It is advisable to check the most favorable span-to-depth ratio by trial designs.

Cross-section
A box girder may have vertical or inclined webs. It is cheaper to manufacture a girder with vertical webs. This
section shape may be the best solution for a narrow road or a single-track railroad. A single narrow closed box
girder can be positioned on the bridge centre line and completed with cantilever brackets. A combination of a
wide deck on a short or medium span bridge favors inclined webs, Figure 1. For instance, a 13 m wide concrete
deck without transverse prestressing requires a width of the box of 6 m at the top. If it were made with vertical
webs, the bottom flange would be
much too wide to be efficient.
Inclined webs reduce the width in a
favorable way.

Fig. 12.
Use of concrete as a lower flange
in case of hogging moment.

20. Design of a steel box girder 305


Dr. A. Romeijn
CT5125: Steel bridges – file <design-box-girder>

Normally the webs are inclined 20 - 35 degrees from the vertical. In many cases inclined webs are chosen for
aesthetic reasons. There are several effects that make wide flanges inefficient. One is shear lag and another is
local buckling of areas in compression. Further, the minimum thickness specified in codes may often make the
flange area excessive. Plate widths of 3,3 m are readily available and some French or German mills produce
wider plates, up to 5 m. (Greater widths are available with thinner plates.). If still wider plates are needed, a
longitudinal weld increases the costs. It is generally preferable to adopt the maximum available width and avoid
the longitudinal weld even if a slightly thicker plate would lead to less stiffening. This advice is valid for the
bottom plate as well as the webs. The best economy is achieved if sections can be fabricated in the full width at
the shop. If the sections can be delivered by boat, the only limitation is the handling equipment. Composite box
girders will frequently be small enough to be shipped in one piece, also by road. Local restrictions for road
transport should be checked. The normal width limit, 2,5 m, may be exceeded if special permits are requested.
The cost of the escort should be checked.

Fig. 13.
Use of corrugated web plates; effective concrete
prestressing.

20.3 Structural Details

Longitudinal stiffeners
Stiffeners are needed on the bottom flange at least at the piers where it is in compression and sometimes also on
the webs. In designing an economical girder, the cost of handling and welding the stiffeners has to be taken into
account. With increasing labor costs the tendency is to have fewer stiffeners and thicker plates. In most cases, the
bottom flange will have a very small effective area if it is not stiffened at the support. An efficient profile is the
cold-formed trapezoidal stiffener. One to two will be sufficient if they are made big enough. If the bridge is to be
erected by launching or cantilevering, it is often necessary to stiffen the bottom flange along the whole girder in
order to resist the hogging moments during erection.

Pier Diaphragms and Intermediate Cross Frames


At the supports, considerable forces from torque and shear have to be transmitted to the bearings. The
recommended solution at piers is a diaphragm, i.e. a steel plate transverse to the girder. The plate is designed to
carry the shear from the torque and is strengthened locally in order to carry the support reactions. The diaphragm
at the pier sections prevents deformation of the section (distortion of the box cross-section). If the bottom flange
is narrow, it may be necessary to put the bearings outside the flange and to provide the webs with external
stiffeners.
In order to prevent cross-section distortion, the girder is provided with intermediate cross frames (see Figure 1).
The webs and bottom flange have transverse stiffeners at these sections. The intermediate transverse stiffeners
are made of flats or strips of plate when cross bracing’s are used. For the intermediate frames, the bracing’s can
be omitted if the rigidity of the cross frames constituted by web and flange stiffeners is enough. The intermediate
transverse stiffeners are made of a T section when bracing’s are not used. To be considered effective against
distortion, cross frames must comply with the stiffness and strength requirements. The stiffness is expressed in
terms of a minimum value for the dimensionless parameter S. This parameter is in effect a ratio between the
distortional stiffness of the box and the distortional stiffness of the cross frame. It may be that the distortional
stresses calculated neglecting all intermediate frames are acceptable, in which case none of the frames need meet
the stiffness criteria.
For the purpose of analysis, box girders may be divided into two groups according to their behavior under
torsional loads. The first group contains girders assumed to have a rigid cross-section, which do not change their

20. Design of a steel box girder 306


Dr. A. Romeijn
CT5125: Steel bridges – file <design-box-girder>

cross sectional shape when rotated about their longitudinal axis. This might be the case for concrete box girders
with relatively thick walls. In other cases, closely spaced diaphragms or bracing’s may be required to satisfy this
condition, like mostly is the case for steel box girders.

Intermediate Transverse Elements between Boxes


The design of the transverse elements
between two longitudinal box girders are
generally subject to important stress
variations under eccentric live loads. Their
design is generally governed by fatigue
considerations. Large, widely spaced
diaphragms may be adapted, figure 14a.
Alternatively, cross girders at 3 to 4m
spacing may be adopted that support a
concrete or orthotropic steel deck as shown
in figure 14b.

Fig. 14. Intermediate transverse elements


between boxes.

Bearings
As a box girder is torsionally rigid, it is possible to use a single bearing at one or more supports and to transmit
the torque to where the foundations are suitable to resist it. This is particularly common if the bridge is highly
curved. Slender columns may support the single bearings.
At each end of the bridge, there are generally two bearings if the bridge consists of a single box girder. In this
case, special attention has to be paid to ensure sufficient distance between the two bearings.
Another consequence of the torsional rigidity is that extra care has to be taken to get the correct support reactions
when there are two bearings at each support. One way of doing this is to let the box rest on jacks with
predetermined loads and to fix the permanent bearing when the jack loads are correct.
If the bearings on the pier are under the diaphragm, care must be taken to ensure that thermal moments do not
lead to longitudinal eccentricity occurring at the bearings. Additional stiffeners may need to be provided.

Corrosion protection
The interior of a box girder is exposed to far less risk of corrosion than the outside. Hence, the interior corrosion
protection can be made simpler or even omitted completely. There is always a possibility of water leaking into
the box, especially if the deck is made of concrete. For this reason, the box should be equipped with a
dehumidifier to keep the air dry. This is an inexpensive precaution.
White painting or very light colors should be used for the interior to facilitate future inspections.

Access holes in diaphragms


Access is usually required through diaphragms, during construction and during service. A manhole 600 mm high
is usually considered as a minimum. Tight corner radii should be avoided; if they are necessary, stress
concentration effects must be carefully considered.
Access into the box (for maintenance) should be either through the end diaphragm or through a hole in the
bottom flange or web. Access through the deck should be avoided, as it is difficult to achieve permanently seals.
Holes in the bottom flange can significantly reduce the flexural strength of the cross section and the entire
structure, especially if they are placed in the vicinity of continuous supports in the effective width region where
high negative moments develop, or in high positive moment regions. Away from high moment regions, normal
stresses due to flexure are small and adding a hole in the bottom flange may be possible (mostly with extra
strengthening because of fatigue criteria).Web access holes can greatly reduce the shear resistance of the web,
which is made of relatively thin steel plates. Therefore, this alternative should only be considered around mid-
span, where low shear forces and torsional moments exist.

20.4 Erection Methods


Box girders may be erected with normal methods such as launching or cantilevering. If the bridge is curved in a
circle, launching works without complication. If the box has an orthotropic deck, it is rigid enough even for

20. Design of a steel box girder 307


Dr. A. Romeijn
CT5125: Steel bridges – file <design-box-girder>

highly curved bridges. However, boxes with composite concrete slabs are normally erected as an open trough.
This open shape is torsionally very soft. The shear centre is unusually far under the centre of gravity, so that the
section will deflect substantially, vertically as well as horizontally, under self-weight, complicating the
launching. Further, the casting of the concrete slab creates additional eccentric load and further deformations and
stresses if the box is curved and open. One solution is to provide the box with horizontal bracing between the top
flanges. The bracing must be designed to avoid interference with the casting of the concrete slab. These
diagonals may be temporary if it is deemed worthwhile to remove them after casting the slab. Some examples of
erection methods are shown.

Fig. 15. Replacement of Moerdijk bridge, 1975-1978.

Box girder parts (22m length) constructed at both


starting points road bridge, transported and lifted by
crane. Temporary supports are used.

Fig. 16. Neckar Vally viaduct , Weitingen,


Germany.

Fig. 17.
Dumbarton bridge.

20. Design of a steel box girder 308


Dr. A. Romeijn
CT5125: Steel bridges – file <design-box-girder>

Fig. 18. Storrow Drive Connector brige.

20.5 Analysis

20.5.1 General
A box girder may be analyzed as a beam subjected to bending, shear and torsion. However, simple beam theory
is not an adequate tool and additional considerations are required, e.g. shear lag, warping and distortion of the
cross-section. The additional stresses caused by cross-section distortion depend largely on the distance between
the cross braces. With a sufficiently small distance, these stresses may be
neglected. National practice varies on this point. The general case of an
eccentric load applied to a box girder is in effect a combination of three
main components: bending, torsion and distortion, figure 20. Warping out
of plane and distortion compromises an internal set of forces, statically in
equilibrium, whose effects depend on the behaviour of the structure
between the point of application and the nearest positions where the box
section is restrained against distortion.

Fig. 19. Normal stress distribution under warping for an I-section beam.

Between points of support, intermediate transverse web stiffeners (intermediate diaphragms or cross frames) may
be provided to develop sufficient shear resistance in a thin web. The cross bracing’s act to restrain cross-section
deformation.

20. Design of a steel box girder 309


Dr. A. Romeijn
CT5125: Steel bridges – file <design-box-girder>

bending

torsion distortion

deformation

warping rotation warping distortion


out of plane in plane out of plane in plane
Fig. 20.
Bending, torsion and distortion components in case of eccentric loading of a boxed girder.

Fig. 21.
Torsion: Typical cross sectional distributions of (a) pure St. Venant’s shear flow, (b) axial stress due to the
restraint of warping, (c) corresponding shear flow.

Fig. 22.
Distortion: Typical cross sectional distributions of (a) distortional axial warping stress, (b) corresponding
distortional shear flow, (c)transverse distortional moment per unit length.

As shown in figures 21 and 22, a state of forces is developed in a (curved) box girder when loaded. The forces
that are developed included bending moments, shear forces, pure (i.e. St. Venant) torsion, warping (i.e. non
uniform) torsional moments, and bi-moments. Torsional moments and bi-moments due to cross-section
distortion also developed. However, providing an adequate number of cross frames can easily reduce distortion-
related effects. When intermediate bracing’s are omitted, special attention should be paid to the design of the
corners of the unbraced cross frames, which should resist the bending moment in the plane of the cross frame.
(The steel box girder works in the same way as pre-stressed concrete box girders). In this case, web and flange T

20. Design of a steel box girder 310


Dr. A. Romeijn
CT5125: Steel bridges – file <design-box-girder>

stiffeners are designed for this purpose. The design of the corners is generally governed by fatigue
considerations. When the intermediate cross frames do not support traffic load directly, they are in general lightly
stressed. The out-of-plane deformation of cross sections subjected to torsion violates the main assumption of the
Bernoulli’s Beam Theory: plane sections remain plane.
If restrained, these out-of-plane deformations create additional normal and shear stresses, which when integrated
over the cross-section yield the bi-moment and warping torsional moment respectively.

Stress calculations
Normal stresses – including the effect of warping – are calculated as follows:

Mx My M M
σ exact = y+ x + ω ω = σ approx + ω ω
Ix Iy Iω Iω

This equation shows that the exact normal stress is generated by the bending moments (M x, M y ) and bi-moment
(M ω) which causes warping. The sum of the first two terms is from classical beam theory, which does not
account for warping, and is frequently referred to as σapprox. The third term is a function of the warping constant,
Iω, and the warping function ω.

Shear stresses - including the effect of warping - are calculated as follows:

Vy V T T T
τ exact = Sx ( s) + x S y ( s) + s + ω Sω ( s ) = τ approx + ω Sω ( s)
tI x tI y 2tAc tI ω tI ω

One important issue is to know the effect of warping on both normal and shear stress. This is achieved by
investigating the ratio between approximate stresses calculated using classical beam theory (i.e. ignoring
warping) and exact stresses, which include the effect of warping.

σ exact − σ approch
WSR − N =
σ approch

τ exact − τ approch
WSR − S =
τ approch

Fig. 23.
Key-points considered for stress calculations.
For most of the bridges constructed (cross-frames
used) for combined loading, total stresses (normal stresses and shear stresses) are underestimated if warping is
not taken into account by an average less than 5% [6]. However, it is recommended that for each bridge a
(global)study on warping stresses be included.

Shear center
Consider the figure below showing a cantilever beam with a transverse force at the tip. Under the action of this
load, the beam may twist as it bends. It is the line of action of the lateral force that is responsible for this bend-
twist coupling. If the line of action of the force passes through the
Shear Center of the beam section, then the beam will only bend
without any twist. Otherwise, twist will accompany bending.

The shear center is in fact the centroid of the internal shear force
system. Depending on the beam's cross-sectional shape along its
length, the location of shear center may vary from section to section.

20. Design of a steel box girder 311


Dr. A. Romeijn
CT5125: Steel bridges – file <design-box-girder>

A line connecting all the shear centers is called the elastic axis of the beam. When a beam is under the action of
a more general lateral load system, then to prevent the beam from twisting, the load must be centered along the
elastic axis of the beam. The two following points facilitate the determination of the shear center location.
1. The shear center always falls on a cross-sectional axis of symmetry.
2. If the cross section contains two axes of symmetry, then the shear center is located at their intersection.
Notice that this is the only case where shear center and centroid coincide

If the cross section contains no axis of symmetry or only one


axis of symmetry, the determination of the exact location of
shear center requires a more detailed analysis.

If the applied shear force does not pass


through the shear center, it will force the beam
to twist as it bends. This eccentricity produces
a torque, that will cause an additional shear
flow and shear stress.

20.5.2 Bending
According to the elementary beam theory, when a box girder is loaded, the longitudinal normal stresses induced
in the flanges are assumed to be uniformly distributed across the flange width. However, in most cases,
particularly in a wide flange, these stresses are non-uniformly distributed due to shear deformations of the flange
plates. This phenomenon, characterized by the fact
that the longitudinal stress on a flange near the web is
much larger than that far from the web, is known as
shear lag.

Fig. 24.
Left: stress distribution without shear deformation.
Right: stress distribution with shear deformation.

The procedure commonly adopted in design is to replace the actual width of flange by an effective breadth
which, when used in conjunction with simple beam theory, leads to a correct estimate of the maximum flange
stress or deflection as required. The effective breadth factor for a (stiffened) flange depends on the geometry of
the box girder, the ratio of stiffener area to plate area, support conditions and loading distribution. Detailed
information on shear leg effects in member design like box girders, is given in prEN 1993-1-5 (Eurocode 1:
Design of steel structures, Part 1-5: Plated structural elements). Some information given by the Eurocode is listed
below.

20. Design of a steel box girder 312


Dr. A. Romeijn
CT5125: Steel bridges – file <design-box-girder>

20. Design of a steel box girder 313


Dr. A. Romeijn
CT5125: Steel bridges – file <design-box-girder>

The distribution of shear stresses in case of a vertical loading is given.

qb,0 = 0
Qy b Q b
qb,1' = qb,0 + ⋅ ⋅( h − hc ) ⋅ tu = y ⋅ ⋅ ( h − hc ) ⋅ tu
Ix 2 Ix 2
qb,2 = 0
Qy Qy
qb,1'' = qb,2 − ⋅ a ⋅ ( h − hc ) ⋅ t u = − ⋅ a ⋅ ( h − hc ) ⋅ tu
Ix Ix

Qy  b 
qb,1''' = qb ,1' − qb,1'' = ⋅  + a ⋅ ( h − hc ) ⋅ tu
Ix  2 
Q 1
qb,3 = qb ,1''' + y ⋅ ( h − hc ) 2 ⋅ ⋅ t w
Ix 2
 b 
= y ⋅ (h − hc ) ⋅   + a  ⋅ tu + ⋅t w ⋅( h −hc ) 
Q 1
Ix  2  2 
Q b Q b
qb,4 = qb,5 + y ⋅ hc ⋅ ⋅ tl = y ⋅ hc ⋅ ⋅ tl
Ix 2 Ix 2
qb,5 = 0

20.5.3 Torsion
If the deflected shape caused by warping is not restricted, only St. Venant shear stresses exist.
If the deflected shape caused by warping is restricted, for example caused by:

20. Design of a steel box girder 314


Dr. A. Romeijn
CT5125: Steel bridges – file <design-box-girder>

- the presence of intermediate supports


- existence of intermediate diaphragms at end supports
- change in torsion along bridge length
additional warping stresses exist.
In general for static analyses the torsional warping stresses in box girders are neglected. However, some care
should be taken with this assumption.

A characteristic feature of thin-walled cross sections is


the large out-of-plane warping displacement that may
occur as a result of torsional (and distortional) loading.
In the absence of any longitudinal restraint, the
distribution of free warping due to constant torque is
identical at every cross section.

Fig. 25.
(a) Free warping.
(b) Restrained warping in a single-cell box section.

In addition to rigid-body rotation, each wall element undergoes a degree of shear deformation.

Pure torsion
The theoretical behavior of a thin walled box section subject to pure torsion is dealt with in many textbooks. For
a single cell box, the torque is resisted by a shear flow (St. Venant), which is constant around the box

b
Ts = 4 ⋅ P ⋅ = 2 ⋅ P ⋅ b
2
Ts = H ⋅ h +V ⋅ b = q s ⋅ b ⋅ h + q s ⋅ h ⋅ b = 2 ⋅ qs ⋅ h ⋅ b
Ts
qs =
2 ⋅ b⋅ h
Ts
q s = τ s ⋅ t ( s) =
2⋅F
Ts = ∫ qs ⋅ rS ( s) ⋅ ds where, rs (s) is distance of shear centre.

Rotation in plane
The shear flow produces shear stresses and strains in the walls and gives rise to a twist per unit length, which is
given by the general expression:

dθ dθ Ts
Ts = G ⋅K s ⋅ or (Bartho – equation) ϑ= =
dz dz G ⋅ K s

with
4 ⋅b2 ⋅ h2
Ks = (torsion constant)
b h b
+ 2⋅ +
tu t w tl

20. Design of a steel box girder 315


Dr. A. Romeijn
CT5125: Steel bridges – file <design-box-girder>

Unrestrained warping out of plane


Pure torsion of a thin walled section will also produce a warping of the cross section, resulting in a torsional
warping displacement.
The general expression for the distribution of a sectorial
coordinate is expressed by:

τs
s s
ws ( s) = ws= 0 ( s ) − ϑ ⋅ ∫ rs ⋅ ds + ∫ ⋅ ds
0 0
G

The first integral determines the level of warping around


the equivalent open section formed by making an
imaginary cut in the closed box at s = 0. It is therefore
evaluated around the entire section.

Fig. 26. Distribution of sectorial coordinate.

The second integral, which applies only to the closed part of the section, takes account of shear deformation in
the various wall elements necessary to restore compatibility of warping displacement around the perimeter.

Restrained warping out of plane


If warping is in any way restrained, axial displacements at each section have the same cross-sectional
distribution, as is shown for pure torsion, but with a varying magnitude along the length of the girder.
A warping stress resultant is formed, generally referred to as a bi-moment (B). This warping torsional moment is
expressed by:

d 2θ
M ω = − E ⋅ Iω ⋅ 2
dz

d 2θ M ω
σ ω = E ⋅ω s ⋅ = ⋅ ωs
dz 2 Iω

This permits the familiar equation for longitudional stress in simple beam theory to be extended to include
torsional warping effects in thin walled sections.

Some practical information on stresses caused by restrained warping out of plane is given in e.g. the Dutch code
NEN 6788. A short summary is given.

3 ⋅ ∆M t
σw =
It ⋅ A
A = 2 ⋅ h ⋅ t3 + 2⋅ b ⋅ t

with b⋅t = b 1 ⋅t1 als b 2 ⋅t2 > b 1 ⋅t1


b⋅t = b 2 ⋅t2 als b 2 ⋅t2 < b 1 ⋅t1
and ∆M t change of torsion moment, in Nmm.
It moment of inertia in mm4
A cross-section area (without stiffeners), in mm2
h girder height, in mm
b girder width, in mm
t plate thickness in mm.

20. Design of a steel box girder 316


Dr. A. Romeijn
CT5125: Steel bridges – file <design-box-girder>

20.5.4 Distortion
Distorsion of box girders occurs when the torsional loads are not distributed to the plates of the cross section in
proportion to a uniform St. Venant shear flow. While torsional warping stresses in box girders may be relatively
small, without proper bracing (cross frames), distortional warping stresses can be quite significant. The
distortional behaviour of box girders can be understood by examining how the transverse force components in
the distortional loads are resisted in the girder. Figure 27 (a) shows the typical distorted shape that results in out-
of-plane bending of the plate components. Figure 27 (b) shows the shears that develop in the through-thickness
direction as a result of the distortion. The distortional loads on the flanges and webs are partially resisted by the
shears that develop in the plates. Out-of-plate bending stresses are induced with the corresponding moments
shown in figure 27 (c).

Fig. 27.
Out-of-plane distortional stresses in
a box girder.

Distortional loads are also partially resisted by the in-plane shears that develop on the cross sections of the
individual plated, as shown in figure 28. The large arrows represent the in-plane shears that resist the distortional
loads that are represented by the small arrows on the girder plates.

Fig. 28.
In-plane distortional warping stresses in
box girders.

The individual plates will experience in-plane bending from these shears, and longitudinal bending stresses may
be induced on the cross section. The longitudinal bending stresses are known as the distortional warping stresses,
and a typical distribution of the warping stress in a box girder is shown in figure 28. Providing internal cross
frames that are spaced along the girder length can significantly reduce cross-sectional distortion. To be effective,
these cross frames must possess sufficient stiffness.

20. Design of a steel box girder 317


Dr. A. Romeijn
CT5125: Steel bridges – file <design-box-girder>

The most common type of internal cross


frame in modern box girder bridges is the K-
frame. Figure 29 shows an internal K- frame
subjected to a concentrated load represented
by the forces H and V.

Fig. 29. Brace forces in cross frame under


distortional loads.

Since no torsional moment result from the


components of the distortional load, V and H must satisfy the equilibium: V(a+b) = 2Hh.
Once V and H are determined, the brace forces can be derived from static equilibrium of the K-frame.
Equilibrium of the K-frame in the horizontal direction leads to S = Dh .
The relationship between the horizontal and vertical components of the diagonal must satisfy the equilibrium of
DH h=DV(a/2), which results in an axial strut force of

a a 2a
S = DH = DV = V = H
2h h a +b
Thus, the axial force in a diagonal of the K-frame is

D= D +D =
2V F aI
+G J
2

=
2 LDK
H 2K
2 2 2
V H h V
h h

Following figure 30, the pure torsional and distortional


components in a vertical and horizontal couple is as
follows:

Fig. 30.
Pure torsional and distortional components in a
vertical and horizontal couple.

External torques generated by horizontal forces acting on the top and bottom plates of the box usually result from
curvature of the girder. The effect of the axial ending stress on the top flange can be modeled as an axial force
that is approx. equal to M/h, where M is the
bending moment and h is the girder depth. The
force M/h acts on the curved segment in the
longitudinal direction as shown in figure 31.

Fig. 31.
Approximation of effect of curvature on girder
flange.

20. Design of a steel box girder 318


Dr. A. Romeijn
CT5125: Steel bridges – file <design-box-girder>

Due to the curvature of the segment, the force resultant of (M/h)dθ is developed on the flange in the lateral
direction. Dividing the lateral force by the length of the segment, ds, a distributed load equal to M.(Rh) is
derived. Due to the nature of bending stresses, the
bottom flange is subjected to the same magnitude of
the lateral load in the opposite direction. As a result,
these two horizontal components form a distributed
torsional load equal to M/R acting on the girder.

Fig. 32. Equivalent torsional loads on a curved box


girder.

Therefore, the effect of the curvature can be approx. analyzed by a straight model in which the flange segments
are subjected to longitudinal forces of M/h as well as external load of M/(Rh) in the lateral direction.
In general the distortional behavior depends on interaction between the two sorts of behavior, the warping and
the transverse distortional bending. The behavior has been demonstrated to be analogous to that of a beam on an
elastic foundation (BEF), with the beam stiffness representing the warping resistance and the elastic foundation
representing the transverse distortional bending resistance.

The differential equation is

d 4κ m
E ⋅ I Dω ⋅ 4
+ K Dω ⋅ κ = T
dz 2

The equation for the distortion stiffness of a box girder is

12 ⋅ (b ⋅ I w ⋅ Iu + b ⋅ Il ⋅ I w + 6 ⋅ h ⋅Iu ⋅ Il ) ⋅ E ⋅ Iw
KDω =
(b 2
⋅ I w2 + 2 ⋅ h ⋅I l ⋅ Iw ⋅ b + 2⋅ b ⋅ I w ⋅ h ⋅ Iu + 3 ⋅ h 2 ⋅ Iu ⋅ Il )
with
E ⋅ tu3
12 ⋅ (1 − µ 2 )
EIu = bending stiffness upper flange =

E ⋅ tw3
12 ⋅ (1 − µ 2 )
EIw = bending stiffness web =

E ⋅ tl3
12 ⋅ (1 − µ 2 )
EIl = bending stiffness lower flange =

For alternative cross-frames, as an example some equations are given.

Intermediate Kd
diaphragms
Plate KD = G ⋅ td ⋅b ⋅ h
Frame 12 ⋅ (b * ⋅ I w ⋅ I u + b* ⋅ I l ⋅ I w + 6 ⋅ h* ⋅ Iu ⋅ Il )⋅ E⋅ I w
KD =
(b *2 2
⋅ I w2 + 2 ⋅ h * ⋅ Il ⋅ Iw ⋅ b + 2 ⋅ b* ⋅ Iw ⋅ h * ⋅ Iu + 3 ⋅ h* ⋅ I u ⋅ I l )
X-shape 2 ⋅ E ⋅ Av ⋅ h2 ⋅ b2
KD = 3
lv
V-shape E ⋅ Av ⋅ h2 ⋅ b2
KD =
2 ⋅ lv
3

20. Design of a steel box girder 319


Dr. A. Romeijn
CT5125: Steel bridges – file <design-box-girder>

Plate
H H
γH = =
G ⋅ A G ⋅ b ⋅ td
V V
γV = =
G ⋅ A G ⋅ h ⋅ td
H H H ⋅h
κ= = = or
G ⋅ A G ⋅ b ⋅td G ⋅ b ⋅td ⋅h

V V V ⋅b
κ= = =
G ⋅ A G ⋅ h ⋅ td G ⋅ h ⋅ t d ⋅ b

TDω
κ=
G ⋅ b ⋅ td ⋅ h
KD = G ⋅ t d ⋅ b ⋅ h

Frame
12 ⋅ (b * ⋅ I w ⋅ I u + b* ⋅ I l ⋅ I w + 6 ⋅ h* ⋅ Iu ⋅ Il )⋅ E⋅ I w
KD =
(b *2 2
⋅ I w2 + 2 ⋅ h * ⋅ Il ⋅ Iw ⋅ b + 2 ⋅ b* ⋅ Iw ⋅ h * ⋅ Iu + 3 ⋅ h* ⋅ I u ⋅ I l )
1
Iu = ⋅ t d ⋅ hu3
12
1
I w = ⋅ t d ⋅ hw3
12
1
I l = ⋅ td ⋅ hl3
12

X-shape

1 1 H 2 ⋅ h2 V 2 ⋅b2
Qv = ⋅ H + V = ⋅
2 2
+
2 2 h2 b2
1 TDω ⋅ ( b + h )
2 2 2 2 2
1 TDω TDω
Qv = ⋅ + 2 = ⋅
2 h2 b 2 h2 ⋅ b 2

⋅ ( b 2 + h2 ) = D ω ⋅ b
T 1 T l
= Dω ⋅
2 b⋅h 2 b⋅h

Qv
uv = ⋅ lb
E ⋅ Av
b Q ⋅b
uv , x = uv ⋅ = v
lb E ⋅ Av
h Qv ⋅ h
uv , x = uv ⋅ =
lb E ⋅ Av

20. Design of a steel box girder 320


Dr. A. Romeijn
CT5125: Steel bridges – file <design-box-girder>

 b h  Qv b 2 + h 2 Qv
2
uv , x u v, y l Q
κ= + =  + ⋅ = ⋅ = b ⋅ v
h b  h b  E ⋅ Av h ⋅ b E ⋅ Av h ⋅ b E ⋅ Av

3
l 1 TDω
κ = 2b 2 ⋅ ⋅
h ⋅ b E ⋅ Av 2

2 ⋅ E ⋅ Av ⋅ h2 ⋅ b2
KD = 3
lb

V-shape

H 2 ⋅ h 2 V 2 ⋅ b2
2
H 
Qv =   + V 2 = +
 2 4⋅ h
2 2
b

TDω
2
TDω
2
TDω 2 ⋅ ( b2 + 4 ⋅ h2 )
Qv = + 2 =
4 ⋅ h2 b 4 ⋅ h 2 ⋅ b2

⋅ ( b 2 + 4 ⋅ h 2 ) = TDω ⋅ b
T 1 l
= Dω ⋅
2 b⋅h b⋅h
Qv
uv = ⋅ lb
E ⋅ Av

b Qv ⋅ b
uv , x = uv ⋅ =
2 ⋅ lb 2 ⋅ E ⋅ Av
h 2 ⋅ Qv ⋅ h
uv , x = 2 ⋅uv ⋅ =
lb E ⋅ Av

 b 2 ⋅ h  Qv b2 + 4 ⋅ h 2 Qv 2 ⋅ lb
2
uv , x u v, y Q
κ= + = +  ⋅ = ⋅ = ⋅ v
h b  2⋅h b  E ⋅ Av 2 ⋅ h ⋅ b E ⋅ Av h ⋅ b E ⋅ Av

2 ⋅ lb
3
T
κ= ⋅ Dω
h ⋅ b E ⋅ Av
2 2

E ⋅ Av ⋅ h2 ⋅ b2
KD =
2 ⋅ lb
3

20. Design of a steel box girder 321


Dr. A. Romeijn
CT5125: Steel bridges – file <design-box-girder>

Example: Brace forces due to distortion of quasi-closed straight box girder

Fig. 33.
Eccentric concrete load on box girder.

Open boxes will require some plane bracing on the open side, to provide torsional stiffness during construction.
It is usual for the deck slab to be concreted in stages and for the steel girders to be unpropped between supports
during this process. Part of the load is thus carried by the steel beam sections alone, part by the composite
sections. A number of separate analyses are therefore required, each representing a different stage.Typically
there are about twice as many stages as spans, since concrete is usually placed alternately in midspan regions and
over supports. Distortional effects should be considered carefully when concreting open sections in stages.
Explanation is given on the following.

A simple span quasi-closed box girder subjected to


distribute top flange loads, as shown in figure 34 (a).
The load is equivalent to an eccentric distributed vertical
load w = 48.2 kN/m with an offset of 0.53m to the
centerline of the girder (shear center vertically). This
loading condition arises in girders with an
unsymmetrical bridge deck, which frequently occurs in
the fascia of bridges. The girder dimensions as well as
the bracing member sizes are shown in figure 34 (b).

Fig. 34.
Quasi-closed straight box girder subjected to eccentric
vertical loading.

The girder length is divided into 16 bracing panels,


each with a length of 3.1 m. K-frames are provided at
every other lateral strut and are spaced at s K = 6.1 m.

Fig. 35.
Quasi-closed box girder (a) cross section and (b)
plan view of truss type.

Solid diaphragms are used at each of the end supports. Other dimensions for the box girder can be obtained from
figure 34, including a = 2.03 m, b = 3.18 m and R = ¥ (straight girder). The corresponding diagonal length is LDK
= 2.50 m.
Substituting these values into results in D = 20.9 kN and S = 8.5 kN respectively.

Predicted values and FEA-results for the K-frame forces in the straight girder with the X-type top lateral truss are
given in figure 36. Only the tensile components of the diagonals and the lateral struts are shown.

20. Design of a steel box girder 322


Dr. A. Romeijn
CT5125: Steel bridges – file <design-box-girder>

Fig. 36. K-frame forces due to eccentric vertical loads.

Example: Brace forces due to distortion of quasi-closed curved continuous box girder
The following case is explained.

A three span curved continuos girder with


concentric loading, as shown in figure 37.

Using the bending moments at the K-frame


locations, the brace forces are calculated with the
equations as given in the previous example with e
= 0. The axial forces in the struts and diagonals at
the interior side of the K-frames are shown in
figure 38 (web inclination is ignored).

Fig. 37.
Curved continuos box girder.

A good agreement exists between FEA results and forces predicted by equations given.
Figure 38 demonstrates that relatively large forces can be generated due to distortion.

When superimposed on forces from bending and torsion,


these forces can become even larger and problems can
develop in these critical bracing’s members since
“typical” sizes are often used in lieu of actual design. In
many situations, relatively small angles that may be
structurally inadequate are used for these bracing
members.

Fig. 38. K-frame forces in curved continuous box


girder.

20. Design of a steel box girder 323


Dr. A. Romeijn
CT5125: Steel bridges – file <design-box-girder>

20.6 Learning from Failures

Structural failures occur as a result of human failures. Moreover, human failures have the inclination to repeat
themselves. During the period 1969-1971, several accidents happened with box girder bridges, all during the
erection stage:
• 1969: bridge across the Danube, Vienna
• 1970: bridge across the Cleddau, Milford Haven, Wales
• 1970: bridge across the Lower Yarra, Melbourne, Australia
• 1971: bridge across the Rhine, Koblenz, Germany.
• 1977: bridge connecting Koror – Babaldaob – New Guinea
These five cases are briefly discussed below.

Vienna
The erection of this bridge proceeded without problems by cantilevering from both sides. The final gap was
closed on a hot summer day. The deformations of the bridge due to temperature expansion are shown in figure
39. During the night an evenly distributed temperature was restored. The bridge straightened, leading to plate
buckling. The buckling was corrected and no collapse occurred.

Fig. 39. Bridge across Danuble 1969 Vienna; buckling of lower flange under compression.

Milford Haven
The midspan of this bridge was erected by cantilevering. With this
method of erection the cross frame above a pier suffers extra loading
due to the cantilevering part. This load causes no problem, provided
that the diaphragm is designed to carry it. This was not the case. The
bridge collapsed, figure 40.

Fig. 40. Cleddau Bridge in Milford, Wales 1970.

20. Design of a steel box girder 324


Dr. A. Romeijn
CT5125: Steel bridges – file <design-box-girder>

Melbourne
The stiffening girder of this cable-stayed bridge consisted of three cells. For erection, the box was divided into
two parts longitudinally, figure 41. On each side of a pier one part of the box was assembled, hoisted to the
correct level and shifted into the correct position to be connected.

Some complications:
• Both parts were asymmetric. Vertical and horizontal deflection
were to be expected due to dead load.
• It is practically impossible to assemble both parts independently in
such a way that the sag is exactly the same. An actual difference of
120 mm was measured.
• The laterally cantilevering top flange is strongly inclined to buckle.

Putting ballast on top of the bridge solved the latter two problems. The
difference in overall deflection disappeared, but buckling of the cantilevering
top flange increased. To solve the problem of the final buckle, some high
strength friction grip bolts were taken out to remove the incompatibility in
flange length, with the disastrous result of passing the ultimate load carrying
resistance. The bridge yielded and collapsed 50 minutes later.

Fig. 41. West Gate Bridge, Melbourne 1970.

Koblenz
Cantilevering was used as the erection method and again a collapse occurred. The failure was due to the
coincidence of three unfavorable aspects, each of which separately would, most probably, not have caused the
collapse.
• Due to welding of the cross weld, a deformation was introduced, increasing the eccentricity of the
compressive stress, Figure 43a.
• A gap of about 460 mm was kept free between the longitudinal stiffeners, so that automatic welding
equipment could pass the stiffeners without stopping, Figure 10b. The buckling length was taken as 460
mm. The effective buckling length was larger, Figure 10c.
• The effect of effective width on the locally unstiffened plate was not taken into account.

20. Design of a steel box girder 325


Dr. A. Romeijn
CT5125: Steel bridges – file <design-box-girder>

Fig. 43.
Rhine River Bridge, Kobelnz 1971.

The accidents in the United Kingdom particularly led to a rigorous investigation program. In 1974, the `Merrison
Rules' were issued, a code giving recommendations on calculation and erection of box girders.

Zeulenroda bridge, East-Germany


Buckling of the lower flange.

Fig. 44. Stausee-Zeulenroda bridge, Germany, 1973.

20. Design of a steel box girder 326


Dr. A. Romeijn
CT5125: Steel bridges – file <design-box-girder>

Koro-Babaldaob bridge
Without any warning, the bridge collapsed. Several explanations have been developed, however, a clear
conclusion about the exact reason doesn’t exist.

Fig. 45. Koror-Baaldaob bridge, Paulau 1977.

REFERENCES

[1] Distortional Loads and Brace Forces in Steel Box Girders, Journal of Strucutral Engineering / Juni 2002.
[2] Distorsie van kokerliggerbruggen, Msc-work M. Oomen, 2003.
[3] Design Guide for Composite Box Girder Bridges, SCI-P-140.
[4] Behaviour of concrete box girder bridges of deformable cross-section, Proc. Instn Civ. Engrs Structs &
Bldgs, 1993, 99 May, 109-122
[5] http://www.ae/msstate.edu/~masoud/Teaching/SA2/
[6] Behavior and design of curved composite box girder bridges, Report Oct. 2002, University of Central
Florida, Orlando, FL 32816-2450.

APPENDIX Examples: analysis

20. Design of a steel box girder 327


Dr. A. Romeijn
CT5125: Steel bridges – file <design-box-girder>

1. Torsion
Opposite torsion moments at fully free girder ends. Box girder dimension: b · h · l = 1950 · 1950 · 10000 mm,
d = 50 mm, E = 210000 N/mm2 and ? = 0,3. Torsion moment: 3,8025 · E+8 Nmm.
4 ⋅ b2 ⋅h 2 4 ⋅ 1950 2 ⋅ 1950 2
Ks = = = 3,7 ⋅ 1011 mm 3
b h b 1950 1950 1950
+ 2⋅ + + 2⋅ +
tu t w tl 50 50 50
dθ Ts 3,8025 ⋅ 108
ϑ= = = = 1,27 ⋅ 10 −8
dz G ⋅ K s 210000
⋅ 3,7 ⋅ 1011
2,6
Rotation girder
dθ 1 1
θ= ⋅ L ⋅ = 1,27 ⋅ 10 −8 ⋅ 10000 ⋅ = 6,35 ⋅ 10−5
dz 2 2

FE-results.

displacement
normal stress (in longitudinal direction)
Because of the unrestrained warping out of plane the normal stresses are negligible (as shown).

2. Torsion
Opposite torsion moments at girder ends. Left end fully clamped and right end fully free.
Girder dimensions identical to first example.

normal stress (in longitudinal direction)


displacement
Because of the restrained warping out of plane, stresses occur (as shown).

20. Design of a steel box girder 328


Dr. A. Romeijn
CT5125: Steel bridges – file <design-box-girder>

3. Torsion
Torsion moment positioned half girder length. Both girder ends pin ended supported and girder dimensions
identical to first example.

displacement
normal stress (in longitudinal direction).
Because of the restrained warping out of plane, mainly stresses occur (as shown) at supports and load input

3. Torsion
Torsion moment uniformly along the whole girder length. Girder pin ended supported at 0 –50 – 100% girder
length and girder dimensions identical to first example.

displacement normal stress (in longitudinal direction)


Because of the restrained warping out of plane, stresses occur (as shown) along the whole girder length

20. Design of a steel box girder 329


Dr. A. Romeijn
CT5125: Steel bridges – file <design-box-girder>

4. Distortion: cross-frame, plate structure


Plate dimension: b · h · t d = 100 · 100 · 10 mm, E = 210000 N/mm2 and ? = 0,3.
Each plate edge loaded uniformly by shear: total edge load F = 1000 N.
Distortion moment: TD? = F · h = F · b = 1000 N · 100 mm. = 100.000 Nmm

load input frame deformation

210000
K D = G ⋅ t d ⋅ b ⋅h = ⋅ 10 ⋅ 100 ⋅ 100 = 8,0769 ⋅ 109 Nmm
2,6
TDω
κ= = = 1,238 · 10-5
G ⋅ b ⋅ td ⋅ h

5. Distortion: cross-frame, frame structure


Frame dimensions: b · h · t d = 100 · 100 · 50 mm, h u = h l = h w = 5 mm, E = 210000 N/mm2 en ? = 0,3. Edge
loading: 750 N.
TD? = F · h = F · b = 750 N · 100 mm. = 75.000 Nmm.

load input. frame deformation

KD =
( )
12 ⋅ b * ⋅ I w ⋅ Iu + b * ⋅ I l ⋅ I w + 6 ⋅ h * ⋅ I u ⋅ Il ⋅ E ⋅ I w
(b *2
⋅ I w + 2 ⋅ h ⋅ I l ⋅ I w ⋅ b + 2 ⋅ b ⋅ Iw ⋅ h ⋅ I u + 3 ⋅ h *2 ⋅ I u ⋅ I l
2 * * *
)
K D = 1,603 ⋅ 107 Nmm

TDω → κ = 4,076 · 10-3


KD =
κ

20. Design of a steel box girder 330


Dr. A. Romeijn
CT5125: Steel bridges – file <design-box-girder>

6. Distortion: girder
Girder pin ended supported at both ends (each end, two bearings) and girder ends stiffened by cross frames
(stiffness 1,696*1013 Nmm).

h= 1875 mm.
b= 2450 mm.
a= 1000 mm.
tu = 150 mm.
tw = 50 mm.
tl = 100 mm.
L= 12700 mm.
E= 210000 N/mm2
?= 0,3

EID? = 3,164·1022 N·mm4 .


7
KD? = 2,970·10 N·mm / mm’.
Mt / 2 = 4800000 N·mm / mm’.
qt = 7680 N / mm’.
girder input as eccentric loading

normal stress displacements

5. Distortion: girder
Identical to example 6, however cross frames located at 0 – 50 – 100% girder length.

normal stress displacements


As an indication:
• the influence of distortion stiffness on stresses is relatively small
• increasing the number of frames by a factor 2 results into a decrease of normal stress by a factor 4.

20. Design of a steel box girder 331


Dr. A. Romeijn

Vous aimerez peut-être aussi