Sunteți pe pagina 1din 73

Lecture 10.

1: Composite Construction - General


OBJECTIVE/SCOPE

To introduce steel-concrete composite members and construction; to explain the composite


action of the two different materials and to show how the structural members are used,
particularly in building construction.

PREREQUISITES

Lecture 7.2: Cross-Section Classification

RELATED LECTURES

All subsequent lectures in Group 10.

SUMMARY

The two complementary materials, structural steel and reinforced concrete, are introduced
and it is shown how composite action is achieved in the case of composite slabs, beams and
columns. The use of composite construction for buildings and bridges is outlined and
illustrated by several typical examples; its main advantages are also illustrated by comparison
with structures of steel and concrete used independently. Attention is drawn to the effect of
this form of construction on other more general problems such as: fire resistance rating, speed
of construction, flexibility and final fitting out.

1. INTRODUCTION

The most important and most frequently encountered combination of construction materials
is that of steel and concrete, with applications in multi-storey commercial buildings and
factories, as well as in bridges. These materials can be used in mixed structural systems, for
example concrete cores encircled by steel tubes, as well as in composite structures where
members consisting of steel and concrete act together compositely.

These essentially different materials are completely compatible and complementary to each
other; they have almost the same thermal expansion; they have an ideal combination of
strengths with the concrete efficient in compression and the steel in tension; concrete also
gives corrosion protection and thermal insulation to the steel at elevated temperatures and
additionally can restrain slender steel sections from local or lateral-torsional buckling.

In multi- storey buildings, structural steelwork is typically used together with concrete; for
example, steel beams with concrete floor slabs. The same applies to road bridges, where
concrete decks are normally preferred. The extent to which the components or parts of a
building structure should embody all steel construction, be constructed entirely in reinforced
concrete, or be of composite construction depends on the circumstances. It is a fact, however,
that engineers are increasingly designing composite and mixed building systems of structural
steel and reinforced concrete to produce more efficient structures when compared to designs
using either material alone. The first two slides give an impression of how and to what extent
composite construction is used for multi-storey buildings: Slide 1 shows a construction site
for a commercial building in London; Slide 2 shows a factory building for the car industry in
Germany.

Slide 1 : Typical composite multi-storey steel-framed building during construction - a


commercial building in London.

Slide 2 : Typical composite multi-storey steel-framed building during execution - a factory


building for the car industry in Germany.
It should be added that the combination of concrete cores, steel frame and composite floor
construction has become the standard construction method for multi-storey commercial
buildings in several countries. Much progress has been made, for example in Japan, where
the structural steel/reinforced concrete frame is the standard system for tall buildings. The
main reason for this preference is that the sections and members shown in Slide 3 are best
suited to resist repeated earthquake loadings, which require a high amount of resistance and
ductility.

Slide 3 : The combination of concrete cores, steel frame and composite floor construction has
become the standard method for multi-storey construction in several countries.

Building with steel and composite elements experienced a renaissance during the 1980's,
resulting in a profusion of new construction concepts and structural details.

Single composite elements, such as isolated beams, columns and slabs (Figure 1), whilst they
are of high quality and resistance, they are also, in many cases, expensive. This is the case
particularly for buildings with small column spacings, floor beam spans well below 9 m and
low loadings. On the other hand, composite floor construction is highly competitive if spans
are increased to 12, 15 or even 20 m. There is, of course, a demand for larger column- free
spans in buildings to facilitate open planning or greater flexibility in office layout, as shown
in Figure 2.
A further important consideration is that the use of rolled steel sections, profiled metal
decking and/or prefabricated composite members speeds up execution. For maximum
efficiency and economy the joints should be cheap to fabricate and straightforward to erect on
site.

Many experts feel that the further development of steel framed buildings depends largely on
the use of composite construction. Unfortunately these two important building materials, steel
and concrete, are promoted by two different industries. Since these industries are in direct
competition with each other, it is sometimes difficult to promote the best use of the two
materials.

Figure 2 shows three examples of the use of composite floor construction comprising steel
beams and concrete slabs, in buildings: Figure 2a shows a typical office building with offices
on both sides of the corridor, the walls of which are defined by the positions of the internal
columns; Figure 2b shows a large span, column- free structure, which allows a high amount of
flexibility; the structure in Figure 2c has a reduced number of columns, with main and
secondary beams.
2. COMPOSITE ACTION IN BEAMS

Composite beams, subject mainly to bending, consist of a steel section acting compositely
with one (or two) flanges of reinforced concrete. The two materials are interconnected by
means of mechanical shear connectors. It is current European practice to achieve this
connection by means of headed studs, semi-automatically welded to the steel flange, see
Slide 4.

Slide 4 : Composite beams, subject mainly to bending, consist of a steel section acting
compositely with one (or two) flanges of reinforced concrete.
Figure 3 shows several composite beam cross-sections in which the wet concrete has been
cast in situ on timber shuttering. For single span beams, sagging bending moments, due to
applied vertical loads, cause tensile forces in the steel section and compression in the
concrete deck thereby making optimum use of each material. Therefore, composite beams,
even with small steel sections, have high stiffness and can carry heavy loads on long spans.

If slip is free to occur at the interface between the steel section and the concrete slab, each
component will act independently, as shown in Figure 4. If slip at the interface is eliminated,
or at least reduced, the slab and the steel member will act together as a composite unit. The
resulting increase in resistance will depend on the extent to which slip is prevented. It should
be noted that Figure 4 refers to the use of headed stud shear connectors. The degree of
interaction depends mainly on the degree of shear connection used.
The following definitions are used to make clear the differences between resistance (strength)
and stiffness properties:

 With regard to resistance, distinction is made between complete and partial shear
connection. The connection is considered to be complete if the re sistance of the
composite beam is decided by the bending resistance, not the horizontal shear
resistance.
 Complete or incomplete interaction between the concrete slab and the steel section
results in a more or less stiff composite beam. Such incomplete interaction arises
when flexible connectors such as headed studs are used and slip (relative
displacement) occurs at the steel-concrete interface.
 The use of composite action has certain advantages. In particular, a composite beam
has greater stiffness and usually a higher load resistance than its non-composite
counterpart, see Figure 5. Consequently, a smaller steel section is usually required.
The result is a saving of material and depth of construction. In turn, the latter leads to
lower storey heights in buildings and lower embankments for bridges.
3. COMPOSITE MEMBERS

3.1 Composite Beams

Figure 3 shows the use of different shapes and types of steel beam (rolled or welded sections)
together with in situ concrete.

Instead of an in situ concrete slab, precast concrete floor or deck units can be used, see Figure
6. Careful detailing and construction practice are needed to ensure adequate containment for
the connectors. Figure 6a shows a system using large prefabricated deck elements with
longitudinal joints. The gaps between the units would be filled with mortar in the final
structure, thereby giving composite action with the beams. Such structural systems were
introduced during the early 1960's. In Germany more than 100 car parks, university, school
and office buildings (see Slide 5) have been built in this way. The use of precast deck units
reduces on-site construction operations and avoids wet trades. The units themselves are cast
on steel formwork in a shop to ensure high quality and small (strict) tolera nces.
Slide 5 : Instead of an in situ concrete slab, precast concrete floor or deck units can be used.

Figure 6b shows thin prefabricated concrete elements, supported by the steel beam flange.
These elements act as permanent formwork when casting the in situ concrete. The transverse
distances between the stud shanks and the edge of the prefabricated concrete element may be
small however, making it difficult to ensure adequate containment for the connectors. The
main reason for the use of these thin plate elements (usually 4-5cm thick) is that they are easy
to handle, and almost as convenient to handle as metal decking.

Figure 6b also shows a partly encased composite beam, the voids of which are filled with
concrete. This type of composite section is often used in parts of Europe today, in order to
enhance the fire resistance rating without additional protection measures. The lower steel
flange remains unprotected.

The usual practice however, in the case of commercial and industrial buildings (see Slide 6),
is to construct the floors using metal decking which incorporates additional embossments or
indentations to provide composite action. This is a very economical way to speed up
construction, and is an important part of modern structural systems. The deck supports the
loads developed before and during concreting and later acts compositely with the in situ
concrete. Steel decking with re-entrant and trapezoidal profiles are typically used, see Figures
7 and 15.
Slide 6 : The usual practice for commercial and industrial buildings is to construct the floors
using metal decking which is embossed to provide composite action.

Composite beams do not need any falsework or timber shuttering. This advantage is
considered in the following section together with two different construction methods,
"propped" and "unpropped".

3.1.1 Propped construction

The efficiency in structural performance will be greatest if it is possible to ensure that the
concrete slab and steel member act compositely at all times. For this purpose, all loads,
including the dead weight of the structure, should be resisted by the composite section. This
requirement can be met by supporting the steel beam until the concrete has hardened. Such
support is known as "propping". The number of temporary supports need not be high;
propping at the quarter-span points and mid-span is generally sufficient. The props are left in
place until the concrete slab has developed adequate resistance.

Different construction methods lead to different stress states, force distributions and
deflections under service conditions. However, composite beams loaded up to failure fail at
the same bending moment (assuming local instability is prevented) irrespective of whether
propped or unpropped construction has been used. Their bending resistance can be easily
calculated by means of rectangular stress blocks as outlined below.

3.1.2 Resistance of section

A typical form of composite construction consists of a slab connected to a series of parallel


steel members. The structural system is therefore essentially a series of interconnected T-
beams with wide, thin concrete flanges as shown in Figure 2. In such a system, the flange
width may not be fully effective in resisting compression due to "shear lag". This
phenomenon, which is taken into account by the well known "effective width" approach, is
explained later.

No account needs usually to be taken of local buckling in the steel section in simply
supported composite beams, since the compression flange is attached to the concrete slab by
shear connectors, and the depth of the web in compression is usually small. In the case of
partial interaction however, the depth of the compressed part of the web is greater. In this
case, therefore, there remains at least theoretically the possib ility that local buckling could
occur in the web of a deep plate girder or in a flange with a wide outstand beyond the shear
connectors.

The dimensions of most steel beam sections in buildings are such that plastic analysis can be
applied to the cross-section of the composite beam. The calculation of the ultimate moment
of resistance is, therefore, an application of the rectangular stress block diagram on the
assumption that the steel sections belong to Class 1 or 2.

3.1.3 Continuous beams and slabs

Many composite beams in buildings are - from the point of view of the static calculation -
continuous beams over simple supports. The concrete slabs are also usually continuous since
they are cast without joints. Continuous beams in comparison with single span bea ms,
therefore, have the following advantages:

 greater load resistance due to the redistribution of bending moments


 greater stiffness
 smaller steel section to withstand the same loading.

On the other hand, the continuity can complicate the design, particula rly in regard to lateral-
torsional and local buckling in negative moment regions. Local buckling of steel can reduce
the bending resistance of the section below the plastic moment, unless certain limitations to
the breadth/thickness ratios of the elements making up the section are met. Based on these
ratios steel sections are grouped in Classes 1 to 4: Class 1 sections allow for global plastic
analysis, using moment redistribution, which gives a very economic design; Class 2 sections
allow for plastic calculation of the moment of resistance but do not permit redistribution. Hot
rolled sections conform to Class 1 or 2 in most cases and when they are used local buckling is
not, therefore, a problem.

Adequately proportioned anti-crack reinforcement should be provided in the concrete slab


over interior supports where joints are not present. If the reinforcing bars have enough
ductility they will increase the bending resistance substantially in these hogging moment
regions.

3.2 Shear Connection

Mechanical connectors are used to develop the composite action between steel beams and
concrete. This connection is provided mainly to resist longitudinal shear, and is referred to as
the "shear connection".

Figure 8 shows several types of shear connectors. They have to fulfil a number of
requirements, as follows:

 they must transfer direct shear at their base.


 they must create a tensile link into the concrete.
 they must be economic to manufacture and fix.
In the industrialised countries the most common connector is the headed stud. It can be
welded semi-automatically (see Slide 4) to the upper flange either directly in the shop or
through thin galvanised steel sheeting on site (see Figure 8a).

Shot fired connectors, as shown in Figure 8b, have been used as an alternative in cases where
metal decking is used and sufficient electrical power is not available on site. These
connectors have the advantage that modified cartridge guns can be used instead of the special
equipment required for complex through-deck welding.

In the case of prefabricated concrete deck units preloaded high strength bolts have sometimes
been used to connect them to the beams, see Figure 8c. This type of connection has been
used, for example, in temporary car parks because the connection can be removed (although
all such existing car parks are in permanent use at the present time).

The behaviour and resistance of headed studs and other connectors are examined by means of
"shear" or "push out" tests. These tests yield load-slip curves such as is shown in Figure 9 for
headed studs. The behaviour is characterised by great stiffness at low loading (under service
conditions) and large deformations at high loadings up to failure. Such ductile behaviour
makes shear force redistribution at the steel-concrete interface possible and allows for partial
shear connection. In addition, headed studs may be spaced uniformly along the beam length
between critical cross-sections.

Composite beams are often designed under the assumption that the unpropped steel beam
supports the weight of the structural steel and wet concrete plus construction loads. It may,
therefore, be decided for reasons of economy to provide only sufficient connectors to develop
enough composite action to support the loads applied afterwards. This approach results in
many less connectors than are required to enable the maximum bending resistance of the
composite beam to be reached. The use of such partial shear connection results in reduced
resistance and stiffness.
Partial shear connection may be unavoidable when a slab is constructed with metal decking.
The number of shear connectors attached to the steel beam may then be limited by the
restriction of being able to place them only in the troughs of the profiled steel sheeting.

3.3 Beam-to-Column Connection

Highly developed connection techniques can be used for connecting together structural steel
members. Economy requires, however, that the joints are economic to fabricate and
straightforward to install on site. Studies have indicated that the cost effectiveness of
composite structures may be improved, if the actual degree of continuity provided by
nominally simple joints is recognised in design.

In composite steel-concrete structures, however, significant additional stiffness and resistance


can be provided simply by placing continuous reinforcing bars in the slab around the
columns, since the single major factor governing the behaviour of joints is the slab action.

This effect can be augmented by a special sequence of construction and concreting, as


follows: during concreting the steel section acts as a single span beam; the beam should be
connected to the steel column by means of double web angles or flange cleats with or without
web angles; after the concrete has hardened (assuming it is without joints as shown in Figure
10c) it is considered as a continuous beam supporting the additional applied loads. By
following this construction sequence, the required bending moment redistribution is not
extensive and plastic rotation can be significantly reduced. In addition the designer can take
the decision whether or not to use shims between the steel compression flange and the
column mainly depending upon the plastic end moment of the joint.
Figure 10 compares simple, rigid and semi-rigid composite joints. The construction detail
without shims, shown in Figure 10c, is consistent with the growing interest in flexibly
connected (semi-rigid) steel frames with simple construction details which speed up
construction. It is proposed that the following performance criteria should be met:

 joints should behave much like a hinge before concreting.


 joints should be stiff and behave elastically up to a predetermined moment value.
 joints must be able to resist the governing plastic moment with adequate plastic
rotation.

Beam-to-column connections in tall buildings demand somewhat different solutions. Until


recently such structural systems employed only simple shear connections between structural
steel and reinforced concrete elements. However, mixed structures should also be considered
which are built by first erecting a frame of light steel columns and deep spandrel beams. The
steel columns are later encased by reinforced concrete.
3.4 Composite Columns

Three different types of composite columns are principally in use, see Figure 11:

 concrete encased steel columns (a)


 concrete filled steel tubes and (c and d)
 rolled section columns partly encased in concrete (b).

In calculating the strength of such columns, full composite interaction without any slip at the
steel-concrete- interface is assumed. Strictly speaking all geometrical and physical non-
linearities of the different materials should be observed. It is only possible, however, to meet
these requirements by using comprehensive numerical methods of analysis and computer
software. The assumed complete interaction enables definition of section properties, and
stiffness and slenderness ratios, for the whole inhomogeneous cross-section. This information
is necessary to determine the load carrying resistance, including slenderness or P--effects.
Eurocode 4 gives simplified design methods for practical use. Instead of more precise
buckling curves, Eurocode 4 [1] has adopted the European buckling curves a,b and c which
were originally established for bare steel columns.

The complete interaction must be ensured by means of mechanical connections. The


connections have to be provided at least at the column ends and where loads or forces are
acting. They should be distributed over the whole cross-section. Such connectors can be
headed studs, top and bottom plates, suitable brackets, vertical gusset plates, shear heads or
other structural means.

Concrete encased columns have the advantage that they meet fire resistance requirements
without any other protection. In addition, they can be easily strengthened by reinforcing bars
in the concrete cover. They do not, however, present an accessible structural steel surface for
later fastenings and attractive surface treatment. In the case of prefabricated encased columns,
the structural steel sections are fabricated in a workshop and include all welds, connection
plates and other necessary attachments. These steel columns (the longest have been up to 30
m long) can then be transported to another workshop, where concreting takes place. After the
concrete encasement has cured the completed columns can be brought to the construction
site.

Concrete filled steel tubes are also in use. The tubes are generally filled with high stre ngth
concrete, with a minimum cube strength of 45 to 55N/mm2 . These strengths, however, are far
below those which have been developed recently in North America.

If the bearing forces from the floor beams are transferred by means of vertical connection
plates, these plates run through the tube and are welded on both sides. This welding ensures
both parts, the steel tube as well as the concrete core, are loaded directly without excessive
slip at the steel-concrete interface. In order to meet the required fire resistance rating, the
concrete core must be longitudinally reinforced. It is impossible, however, to take advantage
of the full column resistance in many cases.

3.5 Partially Encased Steel Sections

Partially encased steel sections, for both beams and columns, are an interesting development
of the last 10 years. The most important feature of such a partially encased section is its
inherent high fire resistance. The fire resistance is due to the fact that the concrete part
prevents the inner steel parts - structural steel as well as reinforcing bars - from heating up
too fast. Figure 12 shows two partly encased composite beams (on the right hand side)
compared with conventional fire protection by means of boards. Slide 7 shows a typical
composite floor construction, where partly encased sections are used; no further fire
protection for beams and slabs is necessary.
Slide 7 : Partially encased steel sections, for both beams and columns, are an interesting
development of the last ten years.
The concrete parts are cast in a workshop or on site before erection. This procedure enables
rapid construction with prefabricated composite members. The concrete between the flanges
should be reinforced by longitudinal bars and stirrups, and should be attached to the web by
stud connectors, welded bars, or bars through holes.

In addition to the enhanced fire resistance, crippling and local buckling of the steel web is
prevented and the resistance of the steel beam against lateral-torsional buckling is
significantly increased. These beams also have greater stiffness under bending and vertical
shear which results in a reduction of final deflection. They look very massive, as can be seen
from Figure 13, and are characterised by their free bottom flange, to which ducts, other
services and plant can be clamped or fastened.

3.6 Composite Slabs

In floor construction, the use of the solid reinforced concrete slab is being replaced more and
more by metal decking, see Figure 14. Modern profiled steel sheeting with additional
indentations or embossments acts as both permanent formwork during concreting and tension
reinforcement after the concrete has hardened. At this final stage the composite slab consists
of a profiled steel sheet and an upper concrete topping which are interconnec ted in such a
manner that horizontal shear forces can be resisted at the steel-concrete interface. Slip
(relative displacements) at the interface must be prevented completely or partly, as should
vertical separation of the steel decking from the concrete topping.
The required composite action can be achieved by various means, see Figure 15. To allow for
the large variety of current and possible future products on the market Eurocode 4 permits the
following methods of achieving shear load:

a) mechanical interlock provided by deformations in the profile (indentations or


embossments).
b) frictional interlock for profiles shaped in a re-entrant form.
c) end anchorage provided by welded studs or shot fired shear connectors.
d) end anchorages by deformation of the ribs at the end of the sheeting in combination
with (b).

The use of profiled steel sheeting undoubtedly speeds up construction. It is also often used
with lightweight concrete to reduce the dead load due to floor construction. In the UK, for
example, this use of lightweight concrete is common practice for commercial buildings.

The composite slabs are supported by steel beams, which normally act compositely with the
concrete slab. The spacing of the beams, and therefore the slab span, depends on the method
of construction, as follows:

 if the beam spacing is about 2,50 m, then no temporary propping is necessary during
concreting of the slab. In this case, the construction stage controls the design of the
metal decking. Due to the short slab span, the stresses in the composite slab in the
final state after the concrete has hardened, are very low. For such floors, trapezoidal
steel sheets with limited horizontal shear resistance and ductility are most often used.
They have the lowest steel weight per square metre of floor area.
 for other floor layouts where the lateral beam spacing is much larger, props are
necessary to support the metal decking during concreting. Due to the longer slab span,
the final composite slab is highly stressed. As a result this final state may govern the
design. In this case the steel sheeting will require good horizontal shear bond
resistance. Re-entrant profiles are often used leading to greater steel weight per square
metre of floor area.

4. COMPOSITE FLOOR CONSTRUCTION

Composite floor construction is essentially an overlay of one-way structural elements. The


slabs span between the secondary or floor beams, which span transversely between the
primary beams. The latter in turn span onto the columns, see Figure 16, Slides 1 and 2. This
set of load paths leads to rectangular grids, with large spans in at least one direction (up to 12,
15 or even 20 m). Up to 15 m, rolled sections are mainly used, while from 12 m upwards
welded plate girders, stub girders or truss girders tend to be more economical.

During the life of a structure changes in use must be expected. Whilst many of these changes
affect service requirements, others will primarily affect layout. The best way to maximise
flexibility of internal planning is to minimise the number of columns. Figure 2 shows typical
examples of ways in which primary beams of larger span, can reduce or eliminate internal
columns. These large span beams may be so deep that services can only be accommodated by
providing holes in the primary webs, see Figure 17. Stiffening around the openings may be
necessary, particularly in the presence of very high vertical shear forces.

Other methods of incorporating services within the structural depth are shown in Figure 18.
One additional alternative is the possibility of tapering the beams near to their ends.
In the case of longer span floors, the designer may need to consider the susceptibility of the
floor structure to vibration. The parameter commonly associated with this effect is the natural
frequency of the floor: the lower the natural frequency, the more the structure may respond
dynamically to occupant- induced vibration.
For this purpose floors (or beams) are normally designed to have a natural frequency not less
than 3Hz, and in the case of floors that may be subject to rhythmic group activities, not less
than 4Hz. An alternative more precise approach is to assess the likely vibrational behaviour
and, taking into account the human reaction to vibration, thereby establish acceptance
criteria.

In summary, composite floor construction used for commercial and other multi-storey
buildings, offers the following main advantages to the designer and client:

 speed and simplicity of construction (metal decking, simple steel connections).


 lighter construction than a traditional concrete building (structural steel and
lightweight concrete, slender structural elements of small dimensions).
 less on site construction (steelwork, prefabricated structural elements).
 small (strict) tolerances achieved by using steel members manufactured under
controlled factory conditions to established quality procedures.

Composite beams are designed using plastic design methods and partial interaction theory,
combining steel and concrete to great effect. To obtain maximum advantage from this form
of construction, planning and design should be integrated from the start. The involvement of
experienced site managers at an early stage will help avoid problems later on. With this
carefully planned approach, different operations such as steel erection, metal decking and
stud welding, concreting, fire protection, cladding, facade work, services and finishing can be
carried out at different floor levels simultaneously.

5. COMPOSITE BRIDGES

Medium span composite bridges are normally constructed from welded, built up, steel plate
girders and a wide reinforced concrete deck, as shown in Figure 19. Box girders, see Figure
20, which look very attractive but which are more expensive, are used less frequently. For the
smaller spans, from 20 up to 35m, rolled steel sections are more popular. They can be used
with a concrete deck slab or embedded in concrete (upper flange and web). Slide 8 gives an
illustration of rolled sections, which can be fabricated curved if required, used in this way.
Slide 8 : Medium span composite bridges are normally constructed from welded, built up,
steel plate girders and a wide reinforced concrete deck.

Since the 1950's, several large span continuous composite highway bridges have been
erected. During the years immediately after World War II, structural steel was very
expensive, and advantage was taken of the light composite cross-section to save material
costs. The sections of today are more compact and simpler, and do not have so many
secondary beams, bracings and stiffeners. This form of structure saves labour costs in the
workshop as well as on the construction site.

Due to the unsymmetrical nature of the cross-section, concrete shrinkage always causes
compression and positive bending in the steel section leading to greater deflections.

In propped construction the compression in the concrete flange due to the self weight of the
beam causes creep deformations. The concrete sheds compression. Stresses and forces are
then redistributed from the concrete flange to the steel beam, and the steel beam, therefore,
has to resist a greater part of the loading. This redistribution also results in increased
deflections.
A simple way to take creep and shrinkage effects into account is to reduce the stiffness of the
concrete by means of appropriate reduction coefficients "n". These n-factors depend not only
on duration and time of loading after concreting, but also on the cross-section properties and
the environmental conditions. It should be noted that this procedure does not apply to beams
in buildings, where less precision is required.

At the ultimate limit state strains due to load are much larger than the strains due to creep and
shrinkage, and the latter can therefore be neglected.

The design considerations for composite bridges are discussed further in Lecture 10.10.

6. CONCLUDING SUMMARY

 Composite construction, particularly that using profiled steel sheeting, allows rapid
construction.
 The weight of steelwork required in composite construction is significantly less than
if the materials were used independently.
 There is no need for expensive falsework and formwork because the steel beam is
able to sustain the self weight of steel and concrete, by itself or with the assistance of
a few temporary props. Timber formwork can be replaced by precast concrete
elements or profiled steel sheeting.
 The aforementioned advantages present a very strong argument for the use of
composite beams in buildings. They are more significant, however, for medium to
long spans than for short spans.
 The main disadvantage of composite construction is the need to provide connectors at
the steel-concrete interface.
 Another minor drawback is that it is somewhat more complicated than other methods
to design and construct. This drawback is particularly relevant to continuous
structures and bridges. However, it is far outweighed by the significant advantages
that can be gained.

7. REFERENCES

[1] Eurocode 4: "Design of Composite Steel and Concrete Structures:" ENV 1994-1-1: Part
1.1: General rules and rules for buildings, CEN (in press).

8. ADDITIONAL READING

1. Bode, H., "Verbundbau, Werner-Verlag", Dusseldorf 1987.


2. Johnson, R.P., "Composite Construction 1 and 2".
3. Hart, F., Henn, W., Sontag, H., "Multi-Storey Buildings in Steel", Second Edition,
Collins, London 1985.
4. Lawson, R.M., "Design of Composite Slabs and Beams with Steel Decking", SCI-
Publication 055, 1989.
5. Bucheli, P., Crisinel, M., "Verbundtrager im Hochbau, Schweizerische Zentalstelle
fur Stahlbau (SZS)", Zurich 1982.
6. Muess, H., "Verbundtrager im Stahlhochbau", Verlag Wilhelm Ernst & Sohn,
Berlin/Munchen/Dusseldorf 1973.
Lecture 10.2: The Behaviour of Beams
OBJECTIVE/SCOPE

To describe the basic behaviour of composite beams including a geometric description of a


typical beam, its construction, and the stress strain relationships that develop under load.

PREREQUISITES

Lecture 7.8.2: Restrained Beams

Lecture 10.1: Composite Construction - General

RELATED LECTURES

Lecture 10.3: Single Span Beams

Lectures 10.4: Continuous Beams

Lectures 10.5: Design for Serviceability

Lectures 10.6: Shear Connection

SUMMARY

Composite beams are described in terms of the steel section, concrete slab and connectors
used in a typical building floor. The material behaviour of each of the components is briefly
reviewed and reference is made to the slender nature of the steel sectio n and the anisotropic
nature of the concrete slab. The structural behaviour of a typical composite beam is
described, in three stages, by reference to the strain and stress in each component part. Firstly
at low loads when full interaction and a linear elastic response occurs; secondly as slip takes
place with increasing load, and finally as the materials reach failure stresses. Propped and
unpropped construction gives rise to different beam behaviour which is described. Partial
interaction is also explained in qualitative terms. The lecture concludes with a summary of
the constraints that the engineer must take into account when designing composite beams.

1. INTRODUCTION

This lecture outlines, in general terms, the behaviour of the most common form of compo site
element - the composite beam. In doing this it will be possible to explain many of the
problems associated with the analysis and design of other elements such as columns and
slabs. The lecture therefore forms a basis on which to build an understanding of composite
behaviour.

A general description of a composite beam is followed by a more detailed discussion the


component parts and their individual structural behaviour. The structural action is described
by reference to the strain, and resulting stress, history of a typical composite beam as it
deforms, under increasing load, to failure.
The way in which composite beams are constructed may alter their resistance to applied
loads. Consequently it is essential to design composite beams for both the construction and
in-service condition. It is also possible to design a beam for "partial connection" so that each
condition is equally critical. A definition of partial connection, and brief reasons for the two-
stage design requirement, is described. Simple single spans are a common form of beam and
their behaviour is explained. The behaviour of continuous spans is also introduced.

Finally, a summary of the design criteria for composite beams is given. These criteria are
covered in more detail in subsequent lectures.

2. COMPONENT BEHAVIOUR

Since a composite beam is formed from three components, see Figure 1, it is necessary to
review the behaviour of each before describing the overall behaviour of the combination.
Under both tension and compression, steel behaves in a linearly elastic fashion until first
yield of the material occurs. Thereafter it deforms in a perfectly plastic manner until strain
hardening occurs. This behaviour is shown diagrammatically in Figure 2a together with the
idealisation of steel behaviour which is assumed for design. In general, most of the steel
section is in tension for simple sagging bending and local buckling of slender sections is not a
problem. However, for continuous beams, significant parts of the steel section are subject to
compression and local buckling has to be considered. This topic will be covered in Lecture
10.4.1 and 10.4.2.

The behaviour of concrete is more complex. Two situations have to be considered. Concrete
in compression follows a non-linear stress/strain curve. This behaviour is shown in Figure 2b
together with the two idealisations used in design. The parabolic stress block is often used in
reinforced concrete design but the rectangular block is normally assumed in composite beam
design. The non-linear material behaviour gives rise to an inelastic response in the structure.
Concrete in tension cracks at very low loads and it is normally assumed, in design, that
concrete has no tensile strength.
The connection behaviour, see Figure 2c, will be covered in detail in Lecture 10.6.1. It is
sufficient, here, to say that it is also non- linear. This behaviour adds to the complexity of
design.
3. DESCRIPTION OF A SIMPLY SUPPORTED COMPOSITE BEAM

3.1 General

Composite beams are formed with a solid, composite or precast concrete slab spanning
between, and connected to, the steel sections. Figure 1 shows a typical layout. The steel parts
are often confusingly referred to as the "beams". In this lecture they are called "steel sections"
to avoid confusion .

The slab usually spans between parallel steel sections and its design is normally dictated by
this transverse action. Consequently the span, depth and concrete grade are determined
separately and are known prior to the beam design.

For non-composite construction, the steel sections alone are designed to carry the load acting
on the floor plus the self weight of the slab, as shown in Figure 3. The steel section is
symmetric about its mid depth and has a neutral axis at this point. The section strains around
this neutral axis and both the outer fibre tensile and compressive stresses are identical. The
stresses () in tension (t) and in compression (c) in the steel section may be evaluated using
simple bending theory.

t and c = Mservice load / Wsteel section

The concrete slab is not connected to the steel section and therefore behaves independently.
As it is generally very weak in longitudinal bending it deforms to the curvature of the steel
section and has its own neutral axis. The bottom surface of the concrete slab is free to slide
over the top flange of the steel section and considerable slip occurs between the two. The
bending resistance of the slab is often so small that it is ignored.
Alternatively, if the concrete slab is connected to the steel section, both act together in
carrying the service load as shown in Figure 4. Slip between the slab and steel section is now
prevented and the connection resists a longitudinal shear force similar in distribution to t he
vertical shear force shown.

The composite section is non-symmetric and shown a single neutral axis often close to the
top flange of the steel section. The tensile and compressive stresses at the outer fibres are
therefore dependent upon the overall moment of inertia (I) of the composite section and their
distance from the single neutral axis.

Assuming that the loading causes elastic deformation the stresses generated in the section
may be determined using simple bending theory. The stresses for the service load condition
may be obtained (Figure 4) from:

t = Mservice load * y1 / Icomposite section

c = Mservice load * y2 / Icomposite section

where

y1 is the distance of the extreme steel fibre from the neutral axis

y2 is the distance of the extreme concrete fibre from the neutral axis

The I value of the composite section is normally several times that of the steel section. It can
therefore be seen that, for a similar load, the extreme fibre stresses generated in the co mposite
section will be much smaller than those generated in the non-composite beam.
This difference also has an effect on the stiffness of the beams which will be discussed in
more detail in Lectures 10.5.1 and 10.5.2.

The stresses developed in the slab as it spans transversely to the length of the beam are
assumed not to affect the longitudinal behaviour. They are generally ignored when designing
the composite beam. However the span of the beam often dictates how much of the slab may
be assumed to help in the longitudinal bending action. This assumption will be covered in
more detail in Lecture 10.3. Here half the transverse span, each side of the steel section, is
assumed to be effective in carrying the longitudinal compression.

The connection between the slab and steel section may be made in many ways. In general it is
formed using a series of discrete mechanical keys. The most common form of connector is
the headed stud which is shown in Figure 1. Lecture 10.6.1 and 10.6.2 cover the detailed
behaviour of this connector and also describe several other types.

It can be seen that composite beams form part of a complex flooring system and it is difficult
to separate the transverse and longitudinal actions of the slab. Figure 1 identifies the typical
beam section which is discussed in the remainder of this lecture.

3.2 Structural Behaviour

The way in which a composite beam behaves under the action of low load, medium load and
the final failure load is best described in stages. The load, the bending moment and shear
force diagrams, deformation, strains and stresses within the section are all shown in
diagrammatic form for the three stages and related to the load deflection response in Figures
4-6.

Stage 1 - Figure 4

For very low loads the steel and concrete behaves in an approximately linear way. The
connection between the two carries very low shear stresses and it is unlikely that appreciable
longitudinal slip will occur. The beam deforms so that the strain distribution at mid span is
linear, as in Figure 4, and the resulting stress is also linear.

It can be seen from the strain diagram that, in this case, the slab must be deep as the neutral
axis lies within the concrete. As a result some of the concrete is in tension. It has been
assumed that this concrete cracks and therefore carries no tensile stress. If the slab was thin it
is possible that the neutral axis would be in the steel and then the area of steel above the axis
would be in compression.

This stage corresponds to the service load situation in the sagging moment region of most
practical composite beams.
Stage 2 - Figure 5

As the load increases the shear stress between the slab and steel section gives rise to
deformation in the connection. This deformation is known as 'slip' and contributes to the
overall deformation of the beam. Figure 5 shows the effect of slip on the strain and stress
distribution. For many composite beams slip is very small and ma y be neglected (exceptions
to this assumption will be covered later in this lecture and in Lecture 10.6.2).

This stage corresponds to the service load stage for that class of composite beams which has
been designed as partially connected. This class of composite beam will be described more
fully later in the lecture and in detail in Lecture 10.6.2.
Stage 3 - Figures 6 and 7
Eventually the load becomes sufficient to cause yield strains in one or more of the materials.
Stage 3a

In the case of yield occurring in the steel, plasticity develops and the stress block develops as
shown in Figure 6. It is normally assumed that, for the ultimate limit state, the plastic stress
block develops such that the whole steel section may eventually reach yield as shown by the
dotted line in Figure 6.

Stage 3b

Concrete is not a plastic material. If strains develop such as to cause overstress it is


potentially possible that explosive brittle failure of the slab would occur. This behaviour
would be similar to the brittle failure expected in an over-reinforced concrete beam. The
volume of concrete in most practical slabs means that it is unlikely that this situation could
ever arise in practice.

With increase in stress within the concrete, induced by increasing strain, the stress block
changes from the triangular shape shown in Figure 5 to the shape shown in Figure 6. For
design this shape is difficult to represent in mathematical form and approximations are used.
These approximations will be covered in more detail in Lecture 10.3. For composite beams
the most common approximation is the rectangular stress block shown by the dashed line in
Figure 6 and in more detail in Figure 2b.

Stage 3c

The remaining components of the composite beam that may fail before the steel yields or the
concrete crushes are the connectors. As the load increases the shear strain, and therefore the
longitudinal shear force between the concrete slab and steel section, increases in proportion.
For a uniformly loaded, single span, composite beam which is assumed to deform in an
elastic manner the longitudinal shear force per m length of the beam (T) between slab and
steel section can be obtained from the expression:

T = V S/I

where S is the first moment of area.

Since the longitudinal shear force is directly proportional to the applied vertic al shear force,
the force on the end connectors is the greatest. For low loads the force acting on a connector
produces elastic deformations. This the slip between the slab and the steel section will be
greatest at the end of the beam. The longitudinal shear and deformation of a typical
composite beam, at this stage of loading, are shown in Figure 7a.

If the load is increased the longitudinal shear force increases, and the load on the end stud
may well cause plastic deformation. A typical load slip relationship for the connectors is
shown in Figure 7. The ductility of the connectors means that the connectors are able to
deform plastically whilst maintaining resistance to longitudinal shear force. Figure 7b shows
the situation when the two end connectors are deforming plastically.

Increasing applied load will produce increasing longitudinal shear and connector
deformation. In consequence, connectors nearer to the beam centreline also begin
sequentially to deform plastically. Failure occurs once all of the connectors have reached
their ultimate resistance as shown in Figure 7c. This sequence of shear load and connector
straining is shown in an exaggerated manner in Figures 7a, b, and c.

This failure pattern is dependent upon the connectors being able to deform p lastically. The
end connector in Figure 5 must be able to deform to a considerable extent before the
connector close to the beam centreline even reaches its ultimate capacity. This requirement
for ductility will be discussed further in Lecture 10.6.1 where it will be shown to dictate beam
span.

It can be seen that the failure of the composite beam is dictated by the resistance of its three
main components. As the elastic interaction of these components is very complex it is normal
to design these sections assuming the stress distribution shown in Figure 2b.

Composite beams designed to fail when the steel yields, the concrete just reaches a failing
strain and all of the connectors deform plastically would appear to be the ideal situation.
There are however several reasons why this situation rarely occurs. The reasons are
investigated below.

3.3 Practical Load Situations

It has been assumed so far that the loading on the beam is uniformly distributed and gives rise
to a parabolic bending moment diagram. This is a common situation but it is also equally
possible to find situations where concentrated loads act on beams.

In the case of uniform loading the maximum bending moment occurs at mid span. This
section is then termed the critical section in bending. The stress block at the critical section is
that described in Figure 6. It results in a longitudinal shear distribution to the shear
connectors shown in Figure 7c. It can be seen that the lo ngitudinal shear developed at the
critical section must be resisted by the connectors between this point and the end of the beam.
It can be deduced that, if the critical section is closer to the beam end, as would be the case
for a single point load close to the support, the number of connectors between this point and
the support needs to increase.

In practice the number of connectors between each load point on a beam subject to multiple
point loads must be determined. This calculation often gives rise to variable spacing of
connectors along the span length.

Point loads may also give rise to high vertical shear force. Although some of the vertical
shear may be carried by the slab and beam flanges, it is common practice to ignore that and
assume all the vertical shear is carried by the web of the steel section.

For continuous beams, discussed later in the lecture, there is a possibility of high shear and
bending occurring together. In this case the moment resistance of the section is reduced. This
aspect has been covered in detail inLecture 7.8.2 and is also discussed in Lecture 10.4.2.

3.4 Creep and Shrinkage

Concrete is subject to two phenomena which alter the strain and therefore the deflection of
the composite beam.
During casting the wet concrete gradually hardens through the process of hydration. This
chemical reaction releases heat causing moisture evaporation which in turn causes the
material to shrink. As the slab is connected to the steel section through the shear connectors,
the concrete shrinkage forces are transmitted into the steel section. These forces cause the
composite beam to deflect. For small spans this deflection can be ignored, but for very large
spans it may be significant and must be taken into account.

Under stress, concrete tends to relax, i.e., to deform plastically under load even when that
load is not close to the ultimate. This phenomenon is known as creep and is of importance in
composite beams. The creep deformation in the concrete gives rise to additional, time
dependent, deflection which must be allowed for in the analysis of the beam at the service
load stage.

3.5 Propped and Unpropped Composite Beams

The geometry of most composite beams is often predetermined by the slab size, as previously
discussed, and by the capability of the steel section to carry the load of wet concrete during
construction. This construction limitation gives rise to two composite beam types, the
propped and the unpropped composite beam.

Consider first the case of the propped beam shown in Figure 8. During construction the steel
section is supported on temporary props. It does not have to resist significant bending
moment and is therefore unstressed and does not deflect. Once the concrete hardens the props
are removed. Each of the component parts of the beam then takes load from the dead weight
of the materials. However, at this stage, the beam is acting as a composite element and its
stiffness and resistance are very much higher than that of the steel section alone. The
deformation due to dead loads is, therefore, small. Any further live loading causes the beam
to deflect. The total stresses present in the beam can be found by summing the stresses due to
dead and live loads.
Consider now the unpropped beam shown in Figure 9. During construction the steel section is
loaded with the dead weight of wet concrete. The steel section is stressed and deforms. The
concrete and the connectors remain largely unstressed, apart from the shrinkage stresses
developed within the hardened concrete. It can be seen, in Figure 9, that the wet concrete
ponds, i.e. the top surface of the concrete remains level and the bottom surface deforms to the
deflected shape of the steel section. The dead load due to the weight of wet concrete is a
substantial proportion of the total load and the stresses developed in the section are often
high.
Additional live loads are carried by the composite sectio n which has almost the same stiffness
as that of the propped beam. The stresses present in the unpropped section can therefore be
obtained by summing the wet concrete stresses and the composite stresses. This calculation
leads to a different stress distribution in the section to that present in the propped composite
beam. However the yield stresses developed in the steel and concrete are the same in both
cases and both unpropped and propped composite beams carry the same ultimate load.

The steel section of an unpropped composite beam often needs to be substantial so that the
weight of wet concrete can be carried. The section is, in fact, often substantially larger than
would be required if the beam had been propped.

The load deflection response of a steel section alone and of a composite beam, both propped
and unpropped, is shown in Figure 10. The strains present and stresses developed are shown
in sequence with the section upon which they act. In the unpropped case the steel section
alone takes the load of wet concrete and the strains due to this wet concrete load are added to
the strains caused by the subsequently applied service loads. The resulting stresses are shown
in the stress block. Whilst the overall deflection of the unpropped beam may be larger tha n
the propped beam at the working load stage this is often not important as the deflection
occurring during construction can be hidden by the finishes.
Despite the drawbacks discussed above, unpropped construction is often preferred for the
following reasons:

 the extra cost involved in providing props.


 the restricted working space available in propped areas.
 the adverse effect on speed of construction.

3.6 Partial Connection

In unpropped construction the size of the steel section is often determined by the weight of
wet concrete, and the size of the slab is determined independently by its transverse span. If
sufficient connectors are provided to transfer the maximum longitudinal force in the steel
section or concrete slab, the resistance of the unpropped composite beam becomes very high.
Indeed composite beams so formed are often capable of carrying several times the required
live load. To avoid providing such excess resistance the partially connected composite
member is used.

It has been assumed so far that the connection will carry all the shear force in the beam up to
the time when the steel section has fully yielded. However, because the resistance of the
unpropped beam is so high, it is often possible to reduce the number of connectors. This
reduction results in a beam where the failure mode would be by connector failure prior to the
steel having fully yielded or the concrete having reached its crushing strength.

Such beams require fewer connectors thereby reducing the overall construction cost. They
are, however, less stiff since fewer connectors allow more slip to occur between the slab and
steel section. Partial connection will be covered more fully in Lecture 10.6.2.

4. CONTINUOUS COMPOSITE BEAMS

Although simply supported beam design is most common there may be situations where use
of continuous beams is appropriate. These beams will be covered in detail in Lecture
10.4.1 and 10.4.2 and only a brief review will be presented here.

The mid span regions of continuous composite beams behave in the same way as the simple
span composite beam. However the support regions display a considerably different
behaviour. This behaviour is shown diagrammatically in Figure 11.
The concrete in the mid span region is generally in compression and the steel in tension. Over
the support this distribution reverses as the moment is now hogging. The concrete cannot
carry significant tensile strains and therefore cracks, leaving only the embedded
reinforcement as effective in resisting moment.

The steel section at the support then has to carry compressive strains throughout a
considerable proportion of its depth. Slender sections are prone to local buckling in this
region and any intervening column section may need to be strengthened to absorb the
compression across its web.
As well as local buckling it is possible that lateral- torsional buckling of the beam may occur
in these regions.

5. CONCLUDING SUMMARY

 Composite beams, subject to sagging moments, fail by yielding of the steel section,
crushing of the concrete slab or shear of the connectors.
 Unpropped composite beams need the steel section to be strong and stiff enough to
carry the weight of wet concrete.
 Partially connected composite beams may be used to ensure economy of shear
connection.
 Continuous composite beams need to be designed to resist both sagging and hogging
bending. The slab reinforcement carries the tensile strain in the hogging region. The
steel section must also be checked for possible buckling.

6. ADDITIONAL READING

1. Book, H., "Verbundbau", Werner Verlag, Dusseldorf, 1987.


2. Johnson, R.P., "Composite Construction 1 and 2".
3. Hart, F., Henn, W., Sontag H., "Multi-storey Buildings in Steel", Second Edition,
Collins, London 1985.
4. Lawson, R.M., "Design of Composite Slabs and Beams with Steel Decking". SCI
Publication 055, 1989.
5. Bucheli, P., Crisinel M., "Verbundtrager im Hochbau", Schweizerische Zentralstelle
fur Stahlbau (S25), Zürich 1982.
6. Huess, H. "Verbundtrager im Stahlhockbau", Verlag Wilhelm Ernst & John Berlin,
Muenchen, D_sseldorf, 1973.
7. Eurocode 4: "Design of Composite Steel and Concrete Structures": ENV1994-1-1:
Part 1.1: General rules and rules for building, CEN (in press).
Lecture 10.3: Single Span Beams
OBJECTIVE/SCOPE

To describe the design of a single span steel-concrete composite beam with full shear
connection, using a plastic design method to determine the internal force distribution at
ultimate limit state; to describe an approximate method to check the deflection at
serviceability limit state.

PREREQUISITES

Lecture 10.2: The Behaviour of Beams

RELATED LECTURES

Lectures 10.4: Continuous Beams

Lectures 10.5: Design for Serviceability

Lectures 10.6: Shear Connection

SUMMARY

This lecture introduces the design criteria for a single span composite beam, concentrating on
the determination of its resistance to positive bending moment, to vertical shear, or to a
combination of both. A plastic design method is used. The conditions for which this method
applies are summarised to show the differences between simply supported and continuous
beams. The design method also assumes that only symmetrical steel sections are used and
that full shear connection between the steel and concrete exists at ultimate limit state. Special
attention is paid to the concrete slab acting as the compression flange of the composite beam.
The effective width and maximum longitudinal shear force of the concrete slab are de fined.
The internal force distribution within the cross-section is described. Formulae based on the
distribution are given which determine the moment and shear resistance of the beam.
Serviceability aspects are also briefly discussed.

1. INTRODUCTION

The object of this lecture is to explain the principles and rules for the design of a simply
supported, i.e. single span, composite steel-concrete beam with full shear connection. Typical
cross-sections of composite beams are shown in Figure 1. For simplicity, only the
symmetrical steel sections 1a, 1c and 1d are considered. The relevant symbols are given in
Figure 2.
For full shear connection the total longitudinal shear resistance of the shear connectors (R q),
distributed between the point of maximum positive bending moment and a simple end
support, must be greater than (or equal to) the lesser of the resistance of the steel beam (R s =
Afy /a) when the plastic neutral axis is in the slab, or the resistance of the concrete flange
(Rc = 0,85 beff hc fck /c) when the plastic neutral axis is in the steel section.

1.1 Ultimate Limit State

The resistance to longitudinal shear (see Figure 3, criterion III) of a shear connection is not
discussed here. An idealised load-slip behaviour of the connector, as illustrated in Figure 4, is
assumed; the type of shear connector which exhibits this behaviour is discussed in Lectures
10.6.1, 10.6.2 and 10.6.3.
This lecture concentrates on the resistance of the beam to moment and vertical shear, which
have maximum values at cross-sections I and II respectively, as shown in Figure 3). Between
these critical cross-sections, each cross-section is subjected to a bending moment and a
vertical shear. This combination is usually only of importance in the case where the loading
includes point or line loads, as shown in Figure 5; here the maximum moment and maximum
vertical shear act together at a critical cross-section adjacent to the point load or line load;
special attention must be paid to this critical cross-section.

In the case of statically determinate beams, such as simply supported single span beams, it is
easy to determine the distribution of bending moments from the equilibrium conditions. To
determine the stress distribution over the cross-section, plastic behaviour is assumed. The
advantage of this method is that the calculation of the resistance is based on the "maximum
moment at failure" condition; this method is also easy to understand and apply.

Steel sections can be classified into 4 classes depending on the local buckling behaviour of
the flange and/or web in compression. In the case of a simply supported single span, plastic
design methods may be used for Class 1 and 2 sections; sections of Class 2 are only allowed
when no rotation capacity is required. These classes are described as follows (see also Figure
6 and Lecture 7.2):

 Class 1: plastic cross-sections which can form a plastic hinge with sufficient rotation
capacity for plastic analysis.
 Class 2: compact cross-sections which can develop the plastic moment of resistance
but have limited rotation capacity.
The steel compression flange, if properly attached to the concrete flange, may be assumed to
be of Class 1.

Table 1 (part of Table 4.2 of Eurocode 4 [1]), classifies steel webs in compression according
to their width to thickness ratios. In a composite beam the compression part of the web, in
positive bending, is always less than half the total depth for a symmetrica l section. A width-
to-thickness ratio less than 83 will, therefore, always be sufficient for a symmetric steel
section in positive bending. Therefore, instability of the web is not critical for the IPE-
sections (according to CEN-EN 19-1986) and the HE-sections (according to CEN-EN 53-
1986).

Because the part in compression is always laterally restrained when the beam is in positive
bending, it is not necessary to check lateral-torsional buckling (see Lecture
10.4.1 and 10.4.2). Other aspects, such as shear buckling are discussed briefly in Section 4.
Web crippling, however, is beyond the scope of this lecture - see Eurocode 3 for further
information [2].

1.2 Serviceability Limit State

For simply supported single spans, the concrete flange is in compression and cracking of the
concrete is not relevant. Only deflections and vibrations are important. Lectures
10.5.1 and 10.5.2 discuss these topics.

2. DESIGN ASPECTS OF THE CONCRETE FLANGE IN COMPRESSION

2.1 Effective width

A typical form of composite construction consists of a slab connected to a series of parallel


steel members. The construction is essentially a series of interconnected T-beams with wide,
thin flanges, as shown in Figure 7(a). In such a system "she ar lag" may cause the flange width
to be not fully effective in resisting compression [3]. This phenomenon can be explained by
reference to a simply supported member, part of whose length is shown on plan in Figure
7(b).
The maximum axial force in the slab is at midspan, while the force at the ends is zero. The
change in longitudinal force is associated with shear in the plane of the slab. The resulting
deformation, shown in Figure 7(b), is inconsistent with simple bending theory, in which
initially plane sections are assumed to remain plane after bending. The edge regions of the
slab are effectively less stiff, and a non-uniform distribution of longitudinal bending stress is
obtained across the section. Simple theory gives an effective value for width, b eff, such that
the area GHJK equals the area ACDEF.

The ratio beff/bv depends not only on the relative dimensions of the system, but also on the
type of loading, the support conditions and the cross-section considered; Figure 7c shows the
effect of the ratio of the beam spacing to span length, b v /L, and the type of loading, on a
simply supported span.

In most codes of practice very simple formulae are given for the calculation of effective
widths, although this may lead to some loss of economy. According to Eurocode 4 [1], for
simply supported beams, the effective width on each side of the steel web should be taken

as , but not greater than half the distance to the next adjacent web, nor greater than the
projection of the cantilever slab for edge beams.

The length lo is the approximate distance between points of zero bending moment. It is equal
to the span for simply supported beams.

A constant effective width may be assumed over the whole of each span. This value may be
taken as the midspan value for a beam.

2.2 Maximum Longitudinal Shear in the Concrete Slab

In the concrete slab, a complex (three-dimensional) force distribution occurs in the region of
the connector. The reason for this behaviour is that bending moments and vertical shear
forces act parallel as well as perpendicular to the beam. It is difficult to find a phys ical design
model for this complex stress distribution, and therefore, most design rules are empirical.
Two design criteria can be identified:

 longitudinal shear in the concrete slab, along the shear planes indicated in Figure 8.
 splitting of the concrete.
It is possible to avoid these failure modes by providing sufficient transverse reinforcement
and choosing the correct distance between the connectors. In some cases, satisfying these
criteria may lead to an increase in concrete slab thickness or resistance.

Longitudinal shear resistances are given in Chapter 6 of Eurocode 4 [1].

If the connectors are welded or shot fired through a continuous profiled steel sheet of a
composite slab, the cross-section of the steel sheet can also be considered as transverse
reinforcement.

3. DESIGN CALCULATION

There are design performance requirements for both the ultimate and serviceability limit
states.

Ultimate Limit State

In designing a composite beam for the ultimate limit state, it is necessary to check the
resistance of the critical cross-sections, and the resistance to longitudinal shear between each
adjacent pair of critical cross-sections (see Figure 3). The forces and moments due to factored
loads are required to be less than the design resistance. This can be expressed by:

Sd  Rd

where

Sd is the design value of an internal force or moment

Rd is the corresponding design value of the resistance

The design value of an internal force or moment, S d, can be determined when the static
system, its geometrical data (when relevant) and the combination of the design values of the
loads are known.

Characteristic values for loadings are given in Eurocode 1: Basis of Design and Actions in
Structures [4]. To determine S d, for example for criteria I of Figure 3, the characteristic
permanent and variable (in this case uniformly distributed) loads must be multiplied by the
corresponding  -factors and combined as follows:

Sd = (l2 /8){GGkj + Q(Qk1 + Q ki)} …………. (1)

which, using the values recommended in Eurocode 4 gives:

Sd = (l2 /8){1,35Gkj + 1,50(Q k1 + Qki)} …………. (2)

where,

Gk,j is the characteristic value of the permanent load

Qk,l is the characteristic value of one of the variable loads


Qk,i is the characteristic value of the other variable loads.

To determine the design resistance, Rd, of members or cross-sections, the design values of the
material strengths and geometrical data (when relevant) are necessary. The design value of a
material property represents its lower characteristic value divided by its corresponding partial
safety factor; the partial factors for material properties (and strengths) are:

Combination Structural Concrete Steel Profiled steel


steel reinforcement sheeting

Fundamental a = 1,1 c = 1,5 s = 1,15 ap = 1,1

Other M values, such as that for the shear connection (studs, friction grip bolts etc.) are given
in Eurocode 4 [1].

The use of these material factors in determining design resistances is shown in Section 4,
Equations (3) to (9), for the case of moment resistance, ie. criterion I of Figure 3.

Serviceability Limit State

In the design of a composite beam for the serviceability limit state, it must be shown that,
under service conditions, the deflections and vibrations do not exceed allowable values and
that cracking of the concrete is limited. The design value of the effect of loads Ed shall be less
than (or equal to) a nominal value C d (or a related function R d):

Ed  Cd or, Ed  Rd

This aspect of design is discussed in greater detail in Lecture 10.5.1.

4. PLASTIC DESIGN METHOD

4.1 Positive Bending Moment

The ultimate load resistance of a simply supported beam is determined by the moment of
resistance of the critical cross-section [5]. The determination of the moment of resistance of
the cross-section is based on the following assumptions:

a) The shear connectors are able to transfer the forces occurring between the steel
and the concrete at failure (full shear connection).
b) No slip occurs between the steel and the concrete (complete interaction).
c) Tension in concrete is neglected.
d) The strains caused by bending are directly proportional to the distance from the
neutral axis; in other words, plane cross-sections remain plane after bending, even
at failure.
e) The relationship between the stress a, and the strain  a of steel is schematically
represented by the diagram shown in Figure 9a.
f) The relation between the stress c and the strain  c of concrete is schematically
represented by the diagram shown in Figure 9b.

Both materials are assumed to behave in a perfectly plastic manner, and therefore, the strains
are not limited. This assumption is similar to that made when calculating the plastic moment
resistance for Class 1 steel sections used independently. The idealised diagram for steel is
shown in Figure 9a. The deviation between the real and the idealised diagram is much smaller
than for concrete as shown in Figure 9b. The use of fck for the maximum stress in the concrete
will clearly result in an unconservative design although in practice the overestimate does not
appear to be very significant. To allow for this overestimate a conservative approximation for
concrete strength (kfck ) is used in design.

Experimental research has proved that the plastic method with k = 0,85, leads to a safe value
for the moment of resistance. This is only true if the upper flange cross-section is less than or
equal to that of the lower flange, as will usually be the case.

Application of these assumptions leads to the stress distributions shown in Figures 10 - 12.
Clearly, the calculation of the moment of resistance M c is dependent on the position of the
neutral axis, which is determined by the relationship between the cross-section of the
concrete slab and the cross-section of the steel beam. Two cases can be identified as follows:
a. the neutral axis is situated in the concrete slab:
1. in the solid part of the composite slab (Rs < Rc ; see Figure 10)
2. in the rib of the composite slab (Rs= Rc)

b. the neutral axis is situated in the steel beam:


1. in the flange of the steel section (Rs > Rc > Rw; see Figure 11)
2. in the web of the steel section (Rs > Rc < Rw; see Figure 13)
The plastic moment resistance, assuming full shear connection and a symmetric steel section,
is expressed in terms of the resistance of various elements of the beam as follows:

Resistance of concrete flange : Rc = beff hc 0,85 fck /c

Resistance of steel flange : Rf = b tf fy /a

Resistance of shear connection : Rq = N Q

Resistance of steel beam : Rs = A fy /a

Resistance of clear web depth : Rv = d t w fy /a

Resistance of overall web depth : Rw = Rs - 2 Rf


where,

A is the area of steel beam

b is the breadth of steel flange

beff is the effective breadth of concrete flange

h is the overall depth of the steel beam

hp is the depth of profiled steel sheet

hc is the depth of concrete flange above upper flange of profiled steel sheet

d is the clear depth of web between fillets

fck is the characteristic cylinder compressive strength of the concrete

Mpl is the plastic moment resistance of steel beam

N is the number of shear connectors in shear span length between two critical cross-sections

Q is the resistance of one shear connector

tf is the thickness of steel flange

tw is the thickness of web

 is

Full shear connection applies when Rq is greater than (or equal to) the lesser of Rc and Rs.

The concrete flange is assumed to be a solid concrete slab, or a composite slab with profiled
steel sheets running perpendicular to the beam. The Equations are conservative for a
composite slab where the profiled steel sheets run parallel to the beam because in the
resistance Rc, the concrete in the ribs is neglected.

For a composite section with full shear connection, where the steel beam has equal flanges,
the plastic moment resistance Mc for positive moments is given by the following:

Case a1:

If the neutral axis is situated in the concrete flange as shown in Figure 10, Rs < Rc and the
positive bending moment of resistance is:

Mpl.Rd = Rs z

where: z = h/2 + hp + hc - x/2


x = (Afy /a) / (beff kfck /c).hc = (Rs /Rc ).hc

Mpl.Rd = Rc (h/2 + hp + hc - Rs.hc/2Rc) ………… (3)

Case a2:

If the neutral axis is situated in the rib of the composite slab, Rs = Rc and Equation (3) can be
rewritten as:

Mpl.Rd = Rs (h + 2hp + hc)/2

or, Mpl.Rd = Rs.h/2+ Rc.(hc/2 + hp ) …………… (4)

Case b1:

If the neutral axis is situated in the steel flange, Rs > Rc > Rw. From equilibrium of normal
forces it can be shown that the axial compression force R in the steel flange (see Figure 11)
is:

Rc + R = Rs - 2 R + R  2 R = Rs - Rc  R = (Rs - Rc )/2

This axial force R is located in the middle of the upper part of the flange, with a depth equal
to: (Rtf)/Rf = (Rs - Rc).tf/2Rf . Therefore, the moment of resistance is equal to the resistance
expressed by the Equation (4) minus (2R)½(Rs - Rc).tf/2Rf equal to (Rs-Rc)2 .tf/4Rf as
illustrated in Figure 11.

This can be written as:

Mpl.Rd = Rs.h/2+ Rc (hc/2 + hp ) - (Rs - Rc)2 .tf/4Rf ……….. (5)

Case b2:

If the neutral axis is in the web of the steel section, Rs > Rc < Rw. In this case, a part of the
web is in compression and, as already discussed, this could influence the classification of the
web.

Webs not fully effective ("non-compact webs") are not treated in this lecture. If the depth to
thickness ratio of the web of a steel section is less than or equal to 83/(1-Rc/Rv )

where  = , it is considered as a compact web and the total depth is effective. The
positive bending resistance is as illustrated in Figure 12:
Mpl.Rd = Rc z + Mpl,N-red.Rd = Rc.(h + 2 hp + hc)/2+ Mpl,N-red.Rd (6)

where:

Mpl,N-red.Rd = plastic moment resistance of the steel beam reduced by a normal force Rc.

According to Eurocode 3 [2] the plastic moment reduced by a normal force for standard
rolled I and H steel sections, can be approximated by:

Mpl,N-red.Rd = 1,11 Mpl.a.Rd (1 - Rc/Rs)  Mpl.Rd ………. (7)


So the resistance can be written as:

Mpl.Rd = Rc.(h + 2 hp + hc)/2 + 1,11 Mpl.a.Rd (1 - Rc/Rs) ……… (8)

Mpl,N-red.Rd can also be written as Mpl.a.Rd - (Rc2 /Rv 2 )(d/4)

In this case the moment of resistance is:

Mpl.Rd = Rc.(h + 2 hp + hc)/2 + Mpl.a.Rd - (Rc2 /Rv 2 )(d/4) ………. (9)

The formulae for the positive moment of resistance values are summarised in Table 2.

4.2 Vertical Shear

Cross-section II of Figure 3 is only subjected to vertical shear. The contribution of the


concrete slab to the resistance to vertical shear is small and difficult to determine and is,
therefore, neglected. Therefore, only the web of the steel section and adjacent parts of the
steel flange are taken into account. The vertical shear resistance, according to Eurocode 3 [2],
is given by:

Vpl,Rd = Av fy /a3) ………… (10)

The shear area Av , for rolled I, H and channel sections loaded parallel to the web, can be
taken as: 1,04 h t w.

In addition, the shear buckling resistance of a steel web must be verified when d/t w>69 for
an unstiffened (and uncased) web. For a simply-supported beam, without intermediate
transverse stiffeners, with full shear connection and subjected to uniformly distributed
loading, Eurocode 4 [1] gives the following simplified rules:

for w 1,5  VRd = Vpl,Rd ……………………………….. (11)

for 1,5 < w< 3,0  VRd = Vpl,Rd (3/ w+ 1/5 w- 1,3) …….. (12)

for 3,0 < w< 4,0  VRd = Vpl,Rd 0,9/ w ……………….. (13)

In accordance with Eurocode 3 the web slenderness w is given by:

w= {fy /3cr}1/2  4 …………… (14)

In practice, an I-section girder has usually a transverse load bearing stiffener at the support,
but no intermediate transverse stiffeners. In such a case the elastic critical shear
resistance cr is given by:
cr = …………………. (15)

If the factored internal force V Sd, is less than Vcr = d t w cr, the shear connectors can be
uniformly distributed; if not, more connectors should be placed near the support.

4.3 Vertical Shear in Combination with Bending Moment

Where the vertical shear VSd exceeds half the plastic shear resistance Vpl.Rd given by Equation
(10) due allowance shall be made for its effect on the plastic moment of resistance. This is of
importance at cross-section (I+II) of Figure 5, where both load effects, vertical shear and
bending, are at a maximum. The following methods are used for such cases:

 If the neutral axis of the composite beam is situated in the concrete slab or in the
flange of the steel section, the previous Equations (3) - (5) can be used where R s is
replaced by the reduced resistance of the steel beam, such that:

Rs,red = Rs - (2V Sd /V Rd - 1) Vpl.Rd for: 0,5 < V Sd/VRd  1 ………….. (16)

In this case, part of the middle of the web of the steel section, is reserved to resist the vertical
shear. The depth of this part of the web is

(2V Sd /VRd - 1) h

providing a depth equal to zero when V Sd /VRd = 0.5 and equal to h where V Sd /VRd = 1,0

 If the neutral axis of the composite beam is situated in the web of the steel section, the
previous Equations (8) and (9) can be used where Mpl.Rd is replaced by a reduced
plastic moment of resistance; according to Eurocode 3 [2] this resistance should be
calculated from the stress distribution given in Figure 13 as briefly discussed below.

The part of the web that is reserved for vertical shear is located in the middle of the web
depth. The section modulus of the web will, therefore, be reduced by:

¼tw ( (2V Sd/VRd - 1) h)2 for 0,5 < V Sd/VRd  1 ………….. (17)

which becomes:

¼tw h2 (1 - (2V Sd /VRd - 1)2 ) ………….. (18)

If the factor  is assumed to be equal to (2V Sd/VRd - 1)2 the section modulus of the web can
also be written as:

¼tw h2 (1 -  )

In other words, it is possible to take into account the vertical shear by reducing the design
yield stress in the web by a factor (1 - ) for the area Av (1 - do /d) where do is the depth of
web neglected when calculating Mpl.Rd. The plastic moment of resistance reduced by vertical
shear can be expressed approximately by:

Mpl,V-red.Rd = {1 - (1 - do /d).A/(2A - Av )}.Mpl.Rd ………….. (19)

These rules are shown in Figure 14 in which Mpl.Rd is the plastic moment of resistance
when  = 1.

5. CONCLUDING SUMMARY

 Plastic analysis of the cross-section is used to determine the positive bending moment
resistance of composite beams (assuming that either Class 1 or 2 sections are used).
 Eurocode 4 gives simple formulae for the effective width of concrete slab acting
compositely with the steel; this "concrete flange" must be detailed to avoid
longitudinal shear and splitting.
 Design of composite beams involves ensuring that the forces and moments due to
factored loads are less than the corresponding design resistance.
 Various expressions for the design positive moment of resistance can be derived;
these depend on the position of the neutral axis relative to the concrete slab, steel
flanges, etc.
 Eurocode 4 gives simplified rules governing the design shear resistance of simply
supported composite beams, with full shear connection, subject to uniformly
distributed loading.
 Where high shear and moment are coincident part of the steel web is reserved to carry
shear, resulting in a decrease in moment resistance.

6. REFERENCES

[1] Eurocode 4: "Design of Composite Steel and Concrete Structures": ENV1994-1-1: Part
1.1: General rules and rules for buildings, CEN (in press).
[2] Eurocode 3: "Design of Steel Structures": ENV1993-1-1: Part 1.1: General rules and rules
for buildings, CEN, 1992.

[3] Dowling, P. J., Knowles, Owens, G., "Structural Steel Design", Steel Construction
Institute, 1988 (UK).

[4] Eurocode 1: "Basis of Design and Actions on Structures", CEN (in preparation).

[5] Stark, J.W.B., van Hove, B. W. E., "Composite Steel and Concrete Beams with Partial
Shear Connection", HERON, TNO-Building and Construction Research/TU-Delft
publication, 2nd quarter 1990 (UK).

7. ADDITIONAL READING

1. Essentials of EC4, prepared by the ECCS-TC 11, "Composite Structures" will be


published early 1993 by the European Convention for Constructional Steelwork,
Brussels.
2. Bode, "Verbundbau, Konstruktion und Berechnung", Werner Verlag, 1987
(Germany).
3. Stark, J. W. B.; "Introduction of Eurocode 4, General Methods of Design of
Composite Construction", IABSE Short Course, Brussels, September 1990, IABSE
Report No. 61, ISBN 3-85748-062-2.
4. Verbundträgen im Hochbau. Theoretische Grundlagen, Beispiele,
Bemessungstabellen, Schweinerische Zentralstelle für Stahlbau, Bericht A3, 1982.
5. Narayanan, R. (ed), Composite Steel Structures. Advances in Design and
Construction. Proceedings of the International Conference, Cardiff, UK, July 1987.
Elsevier Applied Science, London.
Table 1 Maximum width-to-thickness ratios for steel webs
Table 2 Positive moment of resistance values

Rs < Rc

Rs = Rc

Rs > Rc > Rw

Rs > Rc < Rw

OR:

and compact webs

S-ar putea să vă placă și