0% found this document useful (0 votes)
124 views

Ce 3003 Steel Kac Lecture Notes - 2020-2021 - Annotated

This document provides an overview of the CE3003 Steel Design course. The course covers topics such as structural analysis, tension members, compression members, beams, and bolted/welded connections. It includes worked examples and tutorials related to designing structural components. The assessment is a class test worth 50% of the overall course grade. Recommended textbooks are also listed. The contact for the course is Katherine Cashell.

Uploaded by

Manny
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
124 views

Ce 3003 Steel Kac Lecture Notes - 2020-2021 - Annotated

This document provides an overview of the CE3003 Steel Design course. The course covers topics such as structural analysis, tension members, compression members, beams, and bolted/welded connections. It includes worked examples and tutorials related to designing structural components. The assessment is a class test worth 50% of the overall course grade. Recommended textbooks are also listed. The contact for the course is Katherine Cashell.

Uploaded by

Manny
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 109

CE 3003 STEEL DESIGN

K. Cashell
Content of the course:
1. Introduction
2. Basis of design
3. Structural analysis
4. Tension members
5. Local buckling and cross-section classification
6. Compression members
7. Beams
8. Joints – bolted connections
9. Joints – welded connections
10. Introduction to building design

Worked examples:
Example 2.1 Distribution and calculation of design loads
Example 4.1 Tension member design
Example 6.1 Compression member design
Example 7.1 Beam design
Example 8.1 Bolted joint design
Example 9.1 Welded splice design
Example 9.2 Beam-to-column double angle cleat joint

Tutorials
Tutorial 2.1 Completion of Example 2.1 (Design loads)
Tutorial 4.1 Tension member design
Tutorial 6.1 Cross-section compression resistance
Tutorial 6.2 Cross-section and member compression resistance
Tutorial 7.1 Load distribution and beam design
Tutorial 8.1 HSFG bolted joint design
Tutorial 9.1 Bracket design

Assessment:
Class test = 50% of overall CE3003 mark

Recommended texts:
1. Davison, B and Owens, G. W. (Editors). (2012). Steel Designers' Manual. Seventh Edition. The
Steel Construction Institute. Wiley-Blackwell, Oxford.
2. Gardner, L. and Nethercot, D. A. (2011). Designers’ Guide to Eurocode 3: Design of Steel
Buildings. Second Edition. Thomas Telford Publishing, UK.
3. Trahair, N. S., Bradford, M. A., Nethercot, D. A. and Gardner, L. (2008). The behaviour and design
of steel structures to EC3. Taylor and Francis.

Contact:
Katherine Cashell
Room H256, Howell building
[email protected]
1. INTRODUCTION

These notes are for Term 2 of the module CE3003, and cover the analysis and design of steel
structures. The aim of the module is to introduce the principles of preliminary and detailed design of
steelwork and the analysis of stability in steel frames.

Background

Engineering structures are required to support applied loads and transmit these internally through
the structure to the foundations. Applied loading can be in a number of forms (e.g. simple gravity
loading, wind loading, snow loading, traffic loading, dynamic loading and so on). Structural forms can
range from one-dimensional cables to three-dimensional space structures. In practice, the majority
of civil engineering structures, and particularly buildings, are made up of a series of repeated two-
dimensional frames. This course will provide guidance on the general layout of structures and initial
choice of structural system, and will focus on the design of the principal types of structural
components (tension members, compression members and beams), including their
interconnectivity.

Common structural steel products

• Hot-rolled sections - Universal Beams (UB) and Universal Columns (UC), Channels, Angles, Tees,
Hollow Sections, Rails
• Plates, piling
• Cold formed sections and sheeting
• Wires, Ropes, Cables
• Fabricated sections

Relevant British and European Standards

• BS 4 Structural Steel Sections


- Part 1 Hot rolled sections
- Part 2 Hot rolled hollow sections
• BS 5950 Structural use of steelwork in buildings, Parts 1 to 5
• BS 5400 Steel, concrete and composite bridges, Parts 1 to 11
• BS 6399 Loadings for Buildings
• EN 1990 Basis of Design (Eurocode 0)
• EN 1991 Actions on Structures (Eurocode 1)
• EN 1993 Design of Steel Structures (Eurocode 3)
• EN 1994 Design Composite Steel & Concrete Structures (Eurocode 4)
• EN 1998 Design of Structures for Earthquake resistance (Eurocode 8)
Common structural sections

Hot-rolled sections
Light gauge steel sections (cold formed sections)

Fabricated light steel trusses

Fabricated beams
Nominal grades of steel

Hot Rolled Steel

Grade S235 Typical Yield Strength of 235 N/mm2

S275 Typical Yield Strength of 275 N/mm2

S355 Typical Yield Strength of 355 N/mm2

S450 Typical Yield Strength of 450 N/mm2

The key material properties (yield strength fy and ultimate strength fu) for these grades, as defined in
Table 3.1 of Eurocode 3 Part 1.1, are given in Table 1.1. The Young’s modulus E of steel should be
taken as 210000 N/mm2. A typical stress-strain curve for structural steel is given in Figure 1.1.

Table 1.1: Values for yield strength (fy) and ultimate strength (fu)
Steel grade Thickness range Yield strength fy Ultimate strength fu
(mm) (N/mm ) 2
(N/mm2)
S235 t ≤ 40 235 360
40 < t ≤ 80 215 360
S275 t ≤ 40 275 430
40 < t ≤ 80 255 410
S355 t ≤ 40 355 510
40 < t ≤ 80 335 470
S450 t ≤ 40 440 550
40 < t ≤ 80 410 550

Engineering properties of steel

The following properties are important for design:


Figure 1.1: Typical stress-strain curve for structural steel

Advantages and disadvantages of steel

Two of the greatest advantages of steel are:

These are most important where:

• Dead weight is a major factor in design


• Material handling costs are large
• Speed and ease of erection is important
• The job is repetitive
• Site congestion exists
• The environment is hostile
• Transport costs are considerable
• Where it needs to be reused or demolished.
Potential disadvantages of steel:

• Relatively expensive material


• Corrodes unless protected
• Poor behaviour at elevated temperatures unless protected
• Buckling considerations need more attention in design
• Fatigue effects in some special applications

Application of steel in construction:

• Light industrial buildings


• Commercial buildings and hospitals
• Domestic and Educational buildings
• High-rise buildings
• Towers, masts, chimneys, containers
• Offshore construction, power stations
• Bridges
• Temporary supports
2. BASIS OF DESIGN

Choice of structural system

Factors which may influence the choice of a suitable structural system in a steel building include: (i)
the spans involved, (ii) vertical loading applied, (iii) resistance to horizontal loading, (iv) the services
required and (v) the ground conditions.

Design philosophies

Design by Geometrical Ratio

Before mathematics and science were applied to building work, design rules were largely a matter of
experience and tradition. Typically, use was made of geometrical ratios giving limits on what had
been safely built. This approach is still used for some scale buildings; it is also used by designers to
make a first attempt before using more sophisticated techniques. Consequently, geometrical ratios
still make a frequent appearance in current codes to provide initial estimates or as simple
approximations to complicated problems.

Allowable Stress Design

The development of linear elastic structural theories led to the allowable stress format. The
governing principle of design is that the stresses caused by the nominal design loads should not
exceed an allowable stress which is the yield stress reduced by a safety factor. This factor is intended
to cover all expected uncertainties. However, there are severe limitations to this approach,
particularly, its inability to deal with nonlinearities as well as stress and load re-distribution. Progress
in structural theories such as indeterminate analysis and plastic theories provided an incentive to
find a philosophy which could provide a framework for the full range of structural calculations. The
new philosophy that was developed is known as ‘Limit State Design’.

Limit States Design

‘Limit States’ are the various conditions in which a structure would be considered to have failed to
fulfil the purposes for which it was built (e.g. Ultimate Limit State and Serviceability Limit State). The
description of these varies from one code to the other, but a typical list is given in Table 2.1.
Table 2.1: Typical limit states
Ultimate Limit States Serviceability Limit States

There is a general division into collapse (ultimate) and normal working (serviceability) conditions.
The former are those catastrophic states which require a large safety factor in order to reduce their
risk of occurrence to a very low level. The latter are the limits on acceptable behaviour in normal
service. It should be noted that there is always the need for additional or varied criteria for different
structures. Limit states design employs statistical definitions of characteristic loads and resistances,
in conjunction with partial safety factors, to ensure that no limit state is violated.

Types of loads

Loads Fk acting on most structures are specified by appropriate standards or by manufacturers’ data,
etc.

Dead Loads (or Persistent Loads) Gk

These include self-weight of the steel member (usually specified per metre length) as well as other
permanent (stationary) parts of the structure such as concrete floor slabs, finishes, walls, cladding,
etc.). Dead (persistent) loads are calculated either from density or specific weight of the material, or
from manufacturer’s catalogues. Typical values of density and specific weight of commonly used
structural materials are given in Table 2.2. All dead loads have to be used in conjunction with the
appropriate partial safety factor for a given load combination.
Table 2.2: Typical density and specific weight values
Material Density (kg/m3) Specific Weight (kN/m3)

Steel 7850 77
Reinforced Concrete 2420 23.7
Brickwork 2000-2300 20-23
Timber 500-900 5-9

Imposed Loads (or Variable Loads) Qk

These include temporary (or moveable) loads such as snow on roofs, people, furniture, equipment,
moveable partitions, etc. Imposed loads vary with the function of the room or building. Typical
values are given in Table 2.3. All imposed loads are to be used in conjunction with appropriate
partial safety factors for a given load combination.

Table 2.3: Typical imposed load values


Building Usage Imposed Load (kN/m2)

Residential 1.5
Offices 2.5-5.0
Educational 3.0
Theatres 4.0
Workshop 5.0
Storage 2.4 per metre height

Other loads

Other loading may include:

It is important to note that all possible load combinations must be considered to ensure that critical
cases (i.e. those producing maximum moments or forces) are accounted for in design. Load
combinations are discussed in the next section.
Partial safety factors in limit states design

The governing equation for limit state design is given by Equation 2.1, which simply states that the
sum of all loads Fk, multiplied by their associated partial factors γF, must be less than or equal to the
resistance Rk divided by a partial safety factor γM.

∑ (γ
Rk
Fk ) ≤ (2.1)
γM
F

The partial safety factors in limit state design are used to allow for expected variabilities of load, load
combinations, materials, design and detailing procedures, and fabrication and erection procedures.
The values for the partial safety factors are selected to provide an acceptably low probability of
failure (see Figure 2.1).

Figure 2.1: Principle of limit states design

Typically, though not exclusively, γM has a value of 1.0 in structural steelwork calculations (see Table
2.4), whilst γF depends on the uncertainty of load and the probability of simultaneous occurrence of
a given load combination, as shown in the examples of Table 2.5. All values of partial factors are
given as their ‘European’ values, though for implementation in different countries, alternative values
may apply, and reference should be made to each countries’ National Annex.
Table 2.4: Numerical values of partial factors γM for buildings
Partial factor γM Eurocode 3 value
γM0
γM1
γM2

Table 2.5: Typical values of partial load factors for common load combinations
Loading types and combinations

1.35 Dead + 1.5 Imposed


1.35 Dead + 1.5 Wind
1.00 Dead + 1.5 Wind (uplift)
1.35 Dead + 1.5 Imposed + 0.9 Wind
1.35 Dead + 1.05 Imposed + 1.5 Wind

Subscripts and axis conventions in Eurocode 3

Eurocode 3, along with the accompanying suite of structural Eurocodes, makes widespread use of
subscripts in its notation. Although, at first glance, these appear unnecessarily complex, in fact, with
the consistency that exists between codes and the logical way in which the subscripts have been
defined, it is relatively straightforward to become accustomed to their use. For example the
subscripts Ed and Rd refer to ‘design effect’ and ‘design resistance’, so MEd is an applied design
bending moment and MRd is a design bending moment resistance.

For axis convention, the Eurocodes take the x-x axis as along the member length, leaving the y-y and
z-z axes for the cross-section. The y-y axis therefore generally relates to the major cross-section axis
and the z-z to the minor cross-section axis (see examples in Figure 2.2).
Figure 2.2: Axis convention in Eurocodes

Framing schemes

There is a wide variety of steel framing systems that can be adopted for different structural
engineering applications. For building structures, a common solution is to use a steel frame
comprising columns, primary beams and secondary beams, in conjunction with composite or non-
composite floor slabs (see Figure 2.3). In composite construction, a shear connection between the
concrete slab and steel beam is employed to give greater efficiency and reduced beam depth. Other
framing schemes include portal frames, commonly used for low rise industrial buildings, trusses and
space frames, providing light and efficient structural forms, specialist long span systems such as
cellular beams or tapered beams, light steel frames utilising cold-formed elements, and modular
construction, where substantial volumes of complete or partially-complete units are constructed in a
factory environment.
Figure 2.3: Steel frame with composite floor slab

Eurocode 3 gives three possible sets of design assumptions:

1. simple framing – all joints are pinned and the structure is fully braced;
2. continuous framing – all joints are rigid and can transmit all moments and forces;
3. semi-continuous framing – take real stiffness and strength of the joints into account.

Example/Tutorial 2.1 – Calculation and distribution of design loads


Figures 2.4 and 2.5 below shows a simplified structural outline and plan of a two storey building. It is
required to calculate the design loads (using the primary vertical load case of 1.35 Dead + 1.5
Imposed) on the:

• Secondary beams B1 and B2 at both floor levels (roof and first floor)
• Primary beams B3 and B4 at both floor levels (roof and first floor)
• Columns C1 and C2 at both storeys.
Figure 2.4: Simplified structural outline

Figure 2.5: Structural Plan

Assume that there are simple beam-to-beam and beam-to-column connections (so no bending
moment is transmitted between these components).

The direction in which the concrete slab spans is important, since this affects how the load is
transmitted to the steel frame. On the structural plan shown in Figure 2.5, the slab spans vertically
on the page, which means that the load is transmitted as a UDL to the secondary beams, which span
horizontally on the page. The secondary beams then exert end reactions either directly to the
columns (e.g. in the case of B1) or, as point loads, to the primary beams (e.g. B2). The primary beams
exert end reactions to the columns.
Design Data:

Roof: Dead Load = 2.5 kN/m2 (100 mm Reinforced Concrete Slab)

Floor finishes = 1.5 kN/m2

Imposed load = 2.0 kN/m2

Snow Load = 0.5 kN/m2

Perimeter brick parapet = 20 kN/m3 (of height 0.9 m and thickness 10 cm)

First Floor: Dead Load = 3.0 kN/m2 (120 mm Reinforced Concrete Slab)

Floor finishes = 1.5 kN/m2

Imposed load = 5.0 kN/m2

Perimeter brick parapet = 20 kN/m3 (of height 3.0 m and thickness 10 cm)

Assume an unfactored self-weight of 0.4 kN/m for B1 and B2 and of 0.8 kN/m for B3 and B4. This
assumption is dependent upon the anticipated section size. For longer spans and heavier loads, a
higher self-weight should be assumed. Once a beam has been selected, the assumed self-weight can
be compared with the actual self-weight; if there is a significant difference between the two, loads
should be re-evaluated.

Part solution (remainder to be completed as Tutorial 2.1):

Roof beams: (w is the uniformly distributed load per unit length of the beam, and P is the central
point load)

B1: w = 1.35[0.4+(2.5+1.5)×1.1 + (20 ×0.1×0.9)] + 1.5[(2+0.5)×1.1] = 13.04 kN/m

B2: w = 1.35[0.4+(2.5+1.5)×1.1×2] + 1.5[(2+0.5)×1.1×2] = 20.67 kN/m

B3: w = 1.35[0.8+(20×0.1×0.9)] = 3.51 kN/m P = 20.67× (4.5/2) = 46.51 kN

B4: w = 1.35×0.8 = 1.08 kN/m P = 2×20.67×(4.5/2) = 93.02 kN

Floor beams: (w is the uniformly distributed load per unit length of the beam, and P is the central
point load)

B1:

B2:
B3:

B4:

Columns – Top storey:

C1: F1t = RB1 + RB3 = (13.04×4.5/2) + (3.51×4.4/2 + 46.51/2) = 60.32 kN

C2: F2t = 2RB1 + RB4 = 2×(13.04×4.5/2) + (1.08×4.4/2 +93.02/2) = 107.57 kN

Columns – Ground:

C1:

C2:
3. STRUCTURAL ANALYSIS

Before the strength of cross-sections and the stability of members can be checked against the design
requirements, the internal (member) forces and moments within the structure need to be
determined from a global analysis. For simple structures, or by making suitable idealisations (for
initial design purposes) to complex structures, analysis can be performed be means of hand
calculation, though for the detailed design of more complex structures, a computer analysis is often
necessary. Four distinct types of global analysis are possible:

• First order elastic – initial geometry and fully linear material behaviour
• Second order elastic – deformed geometry and fully linear material behaviour
• First order plastic – initial geometry and non-linear material behaviour
• Second order plastic – deformed geometry and non-linear material behaviour.

Typical predictions of load-deformation response for the four types of analysis are shown in Figure
3.1.

Figure 3.1: Prediction of load-deformation response from structural analysis

The simplest form of global structural analysis is a first order elastic analysis where the material
behaviour is assumed to be linear elastic, and any geometric nonlinearities (e.g. P-δ effects) are
ignored. Thus the deformations are proportional to the applied loads. In a second order elastic
analysis, material behaviour remains linear, but geometric nonlinearities including instability effects
are considered. A first order plastic analysis ignores instabilities but takes due account of material
non-linearity (allowing, for example, the formation of plastic hinges). A second order plastic analysis
is the most sophisticated form of global structural analysis and accounts for both material and
geometric nonlinearities.
Elastic analysis is also routinely used to obtain member forces for subsequent use in the member
checks based on the ultimate strength considerations of Section 6 of Eurocode 3. This is well
accepted, can be shown to lead to safe solutions and has the great advantage that superposition of
results may be used when considering different load cases. For certain types of structure e.g. portal
frames, a plastic-hinge form of global analysis may be appropriate; very occasionally for checks on
complex or particularly sensitive configurations a full material and geometrical non-linear approach
may be required.

The choice between a first and a second order analysis should be based upon the flexibility of the
structure; in particular, the extent to which ignoring second order effects might lead to an unsafe
approach due to underestimation of some of the internal forces and moments.

Guidance on the choice between using a first or second order global analysis is given in clause 5.2.1
of Eurocode 3. The clause states that a first order analysis may be used provided that the effects of
deformations (on the internal member forces or moments and on the structural behaviour) are
negligible. This may be assumed to be the case provided Equation 3.1 (Equation 5.1 of Eurocode 3)
is satisfied:

αcr ≥ for elastic analysis (3.1a)

αcr ≥ for plastic analysis (3.1b)

The parameterαcr is the ratio of the elastic critical buckling load for global instability of the structure
Fcr to the design loading on the structure FEd, as given by Equation 3.2.

αcr = Fcr / FEd (3.2)

In many cases experienced engineers will ‘know’ that a first order approach will be satisfactory for
the form of structure under consideration. In case of doubt the check (against Equation 3.1) should,
of course, be made explicitly. Increasingly, standard, commercially available software that includes a
linear elastic frame analysis capability will also provide an option to calculate the elastic critical load
Fcr for the frame.

As an alternative, for portal frames (with shallow roof slopes of less than 26°) and for beam and
column plane frames, for the important sway mode (the form of instability that in most cases is
likely to be associated with the lowest value of Fcr and is therefore likely to be the controlling
influence on the need, or otherwise, for a second order treatment) Equation 3.3 provides an explicit
means for determining αcr using only frame geometry, the applied loads and a first order determined
sway displacement. When utilising Equation 3.3 to determine αcr, Equation 3.1 must be satisfied for
each storey in order for a first order analysis to suffice.

 H Ed  h 
αcr =     (3.3)
 
 VEd   δ H ,Ed 
in which:

HEd is the horizontal reaction at the bottom of the storey due to the horizontal
loads (e.g. wind) and the fictitious horizontal loads

VEd is the total design vertical load on the structure at the level of the bottom of
the storey under consideration

δH,Ed is the horizontal deflection at the top of the storey under consideration
relative to the bottom of the storey, with all horizontal loads (including the
fictitious loads) applied to the structure

h is the storey height.

If Equation 3.1 is not satisfied, second order effects need to be accounted for. This is commonly
achieved, either by amplifying the results (member forces and moments) of a first order analysis by
an amplification factor (Equation 3.4) based on αcr, or by conducting a full second order global
structural analysis. For very flexible frames with αcr ≤ 3, a second order analysis must be conducted.
For regular multi-storey frames, αcr should be calculated for each storey, though it is the base storey
that will normally control.

1
Amplification factor = (3.4)
1 − 1 / α cr

Resistance to sway deformations can be achieved by a variety of means, for example, a diagonal
bracing system (Figure 3.2), rigid connections or a concrete core. In many cases, a combination of
systems may be employed, for example the Swiss Re building in London (Figure 3.3) utilises a
concrete core plus a perimeter grid of diagonally interlocking steel elements.

Figure 3.2: External diagonal bracing system (Sanomatalo Building, Helsinki)


Figure 3.3: Swiss Re Building, London

Eurocode 3 allows for all forms of geometrical and material imperfections (frame and member) to be
included in a second order global analysis and for design to be based directly on the output; such an
approach requires specialist software and is only likely to be used very occasionally in practice, at
least for the foreseeable future. Figure 3.4 shows a global frame imperfection, defined as a sway
mode by the angle ϕ and a (local) member imperfection defined as an initial bow of magnitude e0. A
much more pragmatic treatment separates the effects and considers global (i.e. frame
imperfections) in the global analysis and local (i.e. member imperfections) in the member checks.
Figure 3.4: Global (frame) and local (member) imperfections for structural analysis
4. TENSION MEMBERS

Introduction

A tension member (tie) transmitting a direct pull between two points in a structure is theoretically
the simplest and most efficient structural element. In theory, tension members may be thin and of
small cross-sectional area, in contrast to more stocky compression members, where the cross-
section must be larger in order to resist buckling. In practice, however, the simple theoretical
efficiency is impaired by the end connections required to join tension members to other members in
the frame. The two most common impairments are fastener (bolt) holes and connection eccentricity.
Consequently, end connections need careful attention in design to prevent loss of efficiency.

There are many uses for tension members (some of which are shown in Figure 4.1), such as:

• Tension chords and internal ties in trusses and lattice girders in buildings and bridges
• Tension bracing members in buildings
• Hangers in suspended structures
• Main cables and deck suspension cables in suspension and cable-stayed bridges.

The main types of tension members are listed below, and are illustrated in Figure 4.2.

• Open and closed single rolled sections (e.g. angles, tees, channels & hollow sections); used for
tension members in light trusses and for bracing.
• Compound sections consisting of double angles or channels; specified when loads are higher
than can be carried by a single section; should be interconnected at intervals.
• Heavy rolled and heavy compound sections of built-up H- and box sections; built-up sections
should be tied or battened at intervals (for vibration reduction, load distribution and rigidity);
used in bridges and heavy lattice girders.
• Bars and flats; low stiffness; may sag under own weight and may vibrate under wind or moving
loads.
• Ropes and cables; constructed from high strength wires; provide long and very strong
continuous tension members; used in many applications such as in suspension and cable-stayed
bridges.
End connections and splices

End connections create problems that affect the design of tension members and often result in loss
of efficiency. In bolted joints, the member strength is reduced, firstly by the bolt hole, secondly
because any (non-connected) outstanding leg is not fully effective and there is a moment due to
eccentricity in the connection. The welded joint for a single angle has a similar eccentricity problem
(Figure 4.3). Full strength joints where the member is fully effective may be made by welding.
However, for practical and economic reasons this is not used in many cases. Bars have threaded
ends and may be connected directly through a nut. The strength is determined by the net tensile
area at the thread. Ropes are connected through end sockets or terminals.

Splices are needed to connect long members that are constructed from parts for ease of fabrication
and transport, e.g. tension chords in trusses. Splices for open or closed sections may be bolted or
welded.

Design for axial tension

Under direct tension, rolled sections behave similarly to tensile test specimens, as shown in Figure
4.4 (typical stress-strain curves for structural steel and wire ropes).

For a straight member subjected to direct tension:

F
Tensile stress, ft = (axial tensile stress, uniform across cross-section)
A

FL
Elongation, δL = (in the linear elastic range, i.e. prior to yielding)
EA

Load at yield, Fy = fyA (= load at failure, neglecting strain hardening)

Failure modes

There are two major failure modes for members under direct tension:

1. Member yielding - failure is defined as the point at which the applied load reaches the yield limit
where large spread of plastic flow of deformation is found throughout the member; or in other
words, the member fails to keep up its own configuration as designed. This is a ductile failure
mode (displays deformation prior to failure).
2. Local Fracture – typically occurs around bolt holes etc. where there is reduced area and hence
stress concentrations can build up.
Stress concentrations

Stress concentrations occur at holes and changes in cross-sections in members. They are not usually
important in ductile materials, but can be the cause of failure due to fatigue or brittle fracture in
certain conditions. A hole in a flat member will increase the stress locally on the net section by a
factor of 2 to 3 depending on the ratio of hole diameter to net plate width.

Design for axially loaded tension members according to Eurocode 3

Tension Capacity (clause 6.2.3)

For all tension members, it must be shown that the design value of the tensile force NEd (i.e. the total
factored tensile load in the member) is less than the design tensile resistance of the member Nt,Rd.

In Eurocode 3, similarly to BS 5950 Part 1, design tensile resistance Nt,Rd is limited either by yielding of
the gross cross-section Npl,Rd (to prevent excessive deformation of the member) or ultimate failure of
the net cross-section Nu,Rd (at holes for fasteners), whichever is the lesser.

The gross area of a cross-section utilises nominal dimensions, with no reduction for fastener holes,
but allowance should be made for larger openings, such as those for services. Note that Eurocode 3
uses the generic term ‘fasteners’ to cover bolts, rivets and pins.

In general, the net area of the cross-section is taken as the gross area less appropriate deductions for
fastener holes and other openings. For a non-staggered arrangement of fasteners, for example as
shown in Figure 4.5, the total area to be deducted should be taken as the sum of the sectional areas
of the holes on any line (A-A) perpendicular to the member axis that passes through the centreline of
the holes.

Figure 4.5: Non-staggered arrangement of fasteners


For a staggered arrangement of fasteners, for example as shown in Figure 4.6, the total area to be
deducted should be taken as the greater of:

1. the maximum sum of the sectional areas of the holes on any line (A-A) perpendicular to the
member axis
 s2 
2. t  nd 0 −

∑  measured on any diagonal or zig-zag line (A-B)
4p 

where:

s is the staggered pitch of two consecutive holes (see Figure 4.6).


p is the spacing of the centres of the same two holes measured perpendicular to the member
axis (see Figure 4.6).
n is the number of holes extending in any diagonal or zig-zag line progressively across the
section.
d0 is the diameter of hole.

Figure 4.6: Staggered arrangement of fasteners

The Eurocode 3 design expression for yielding of the gross cross-section (plastic resistance) is
therefore given as:

Afy
Npl,Rd = (4.1)
γ M0

And for the ultimate resistance of the net cross-section, concentrically loaded (defined in clause
6.2.2.2), the Eurocode 3 design expression is:

0.9 A net f u
Nu,Rd = (4.2)
γ M2

The design tensile resistance is taken as the smaller of the above two results. For ductility (capacity
design), the design plastic resistance of the gross cross-section should be less than the design
ultimate resistance of the net cross-section. The values for the partial factors are γM0 = 1.0 and
γM2 = 1.25.

Eccentricity of end connections (clause 3.10.3 of EN 1993-1-8)

Moment arising from eccentric connections must be accounted for (by designing for combined
bending and tension). However, simplified design rules are given in Eurocode 3 Part 1.8 for the effects
of combined tension and bending caused by the eccentric load introduced into the member by the
end connection. These rules only apply to unsymmetrical members and symmetrical members that
are connected unsymmetrically, such as angles connected through one leg.

Single angles in tension connected by a single row of bolts through one leg, may be treated as
concentrically loaded, but with an effective net section, to give the design ultimate tensile resistance
as defined by Equations 4.3 to 4.5 (from clause 3.10.2(2) of EN 1993-1-8).

2.0(e 2 − 0.5d 0 ) tf u
With 1 bolt: Nu,Rd = (4.3)
γ M2

β 2 A net f u
With 2 bolts: Nu,Rd = (4.4)
γ M2

β 3 A net f u
With 3 or more bolts: Nu,Rd = (4.5)
γ M2

where:

β2 and β3 reduction factors dependent upon the bolt spacing p1, as defined in Table 4.1.

Anet is the net area of the angle. For an unequal angle connected by its smaller leg,

Anet should be taken as the net section of an equivalent equal angle of leg length
equal to the smaller leg of the unequal angle.

Other symbols are defined by Figure 4.7.

Table 4.1: Reduction factors β2 and β3


Pitch p1 ≤ 2.5d0 ≥ 5.0d0

β2 (for 2 bolts) 0.4 0.7


β3 (for 3 or more bolts) 0.5 0.7
Note: For intermediate values of pitch p1 values of β may be determined by
linear interpolation. d0 is the bolt hole diameter.
In the case of welded end connections, for an equal angle, or an unequal angle connected by its larger
leg, the eccentricity may be neglected, and the effective area may be taken as equal to the gross area
(clause 4.13(2) of EN 1993-1-8).

If the back-to-back angles are connected to both sides of the gusset, the centroid of the pair of angles
is concentric with the gusset plate and there is no reduction (due to eccentricity) in effectiveness.

Figure 4.7: Definitions for e1, e2, p1 and d0

Serviceability, corrosion and fatigue

As tension members transfer load very effectively they tend to have relatively small cross-sectional
areas. This makes ties susceptible to excessive elongation under axial load – which may allow
relatively large lateral deflections under self-weight. For this reason, good practice limits the
slenderness of tension members to around 300 for principal members and 400 for secondary
members. The slenderness is defined to be L/iz where 𝑖𝑖𝑧𝑧 2 = 𝐼𝐼𝑧𝑧 ⁄𝐴𝐴 and L is the length of the element,
iz is the minimum radius of gyration, lz is the minimum second moment of area of the section and A
is the area of the cross-section. It is also important to consider material loss due to corrosion.

Example/Tutorial 4.1 - Tension member design


Example 4.1 demonstrates the calculations required (to Eurocode 3) to check the resistance against
yielding of the gross cross-sections and ultimate failure of the net cross-section for a welded and a
bolted unequal angle arrangement.
5. LOCAL BUCKLING AND CROSS-SECTION CLASSIFICATION

For efficiency, structural members are generally composed of relatively thin elements (i.e.
thicknesses substantially less than other cross-sectional dimensions), for example, I-sections and
hollow sections, as opposed to solid sections. Although favourable in terms of overall structural
efficiency, the slender nature of these thin elements results in susceptibility to local instabilities
(buckling) under compressive stress, which must be considered in design. Clearly, neither local
buckling (nor member buckling) occurs under tensile stress.

Local buckling is so-called because the wavelengths of the buckles are of the same order of
magnitude as the cross-sectional dimensions, so are therefore localised when compared to the
overall (length) dimensions of the structural component. An example of local buckling is shown in
Figure 5.1.

Figure 5.1: Local buckling in structural components

Determining the resistance (strength) of structural steel components requires the designer to consider
firstly the cross-sectional behaviour and secondly the overall member behaviour. Whether in the
elastic or inelastic material range, cross-sectional resistance and rotation capacity are limited by the
effects of local buckling. Eurocode 3 (and BS 5950) account for the effects of local buckling through
cross-section classification.
In Eurocode 3, cross-sections are placed into one of four behavioural classes depending upon the
material yield strength, the width-to-thickness ratios of the individual compression parts (e.g. webs
and flanges) within the cross-section, and the loading arrangement. The classifications from BS 5950
of plastic, compact, semi-compact and slender are replaced in Eurocode 3 with Class 1, Class 2, Class
3 and Class 4, respectively.

Parameters affecting local buckling

The factors that affect local buckling (and therefore the cross-section classification) are as given below.
Definitions of dimensions and axes for different cross-section types are given by Figure 5.2.

1. Width/thickness ratios of plate components


2. Element support conditions (e.g. internal or outstand elements)
3.
4.

Cross-section classification

In Eurocode 3 classification of a cross-section is made by comparing the actual width-to-thickness


ratios of the plate elements with a set of limiting values for each of the four classes, given in Table 5.2
of the notes (also Table 5.2 of EN 1993-1-1). A plate element is Class 4 (slender) if it fails to meet the
limiting values for a Class 3 element. The classification of the overall cross-section is taken as the least
favourable of the constituent elements (for example, a cross-section with a Class 3 flange and Class 1
web has an overall classification of Class 3). Cross-section classification must always be carried out
before other design checks because different design rules apply to members with different cross-
section classes. If a cross-section is Class 4 (slender) a reduced (effective) cross-sectional area must
be calculated in accordance with Eurocode 3 Part 1.5 to account for the effects of local buckling.
Figure 5.2: Dimensions and axes of sections in Eurocode 3

For a beam subjected to pure bending moment (i.e. equal and opposite end moments M), Figure 5.3
shows the four different moment-rotation (M-θ) relationships for the four different classes of cross-
section.
Figure 5.3: Four behavioural classes of cross-section

The Eurocode 3 definitions of the four classes are as follows (clause 5.5.2(1)):

• Class 1 cross-sections are those which can form a plastic hinge with the rotation capacity
required from plastic analysis without reduction of the resistance.
• Class 2 cross-sections are those which can develop their plastic moment resistance, but have
limited rotation capacity because of local buckling.
• Class 3 cross-sections are those in which the elastically calculated stress in the extreme
compression fibre of the steel member assuming an elastic distribution of stresses can reach the
yield strength, but local buckling is liable to prevent development of the plastic moment
resistance.
• Class 4 cross-sections are those in which local buckling will occur before the attainment of yield
stress in one or more parts of the cross-section.

Class 1 cross-sections are fully effective under pure compression and are capable of reaching and
maintaining their full plastic moment Mpl in bending (and may therefore be used in plastic design).
Class 2 cross-sections have a somewhat lower deformation capacity, but are also fully effective in
pure compression and are capable of reaching their full plastic moment Mpl in bending. Class 3
cross-sections are fully effective in pure compression, but local buckling prevents attainment of the
full plastic moment in bending; bending moment resistance is therefore limited to the (elastic) yield
moment Mel. For Class 4 cross-sections, local buckling occurs in the elastic range. An effective cross-
section is therefore defined based on the width-to-thickness ratios of individual plate elements, and
this is used to determine the cross-sectional resistance Meff. In hot-rolled design the majority of
standard cross-sections will be Class 1, 2 or 3, where resistances may be based on gross section
properties obtained from section tables (See Appendix A of these notes). Effective width
formulations are not contained in EN 1993-1-1 (Part 1.1 of Eurocode 3), but may instead be found in
EN 1993-1-5 (Part 1.5 of Eurocode 3). Cross-sectional resistances in compression and in bending for
the four behavioural classes is summarised in Table 5.1.

Table 5.1: Cross-sectional resistances for the four behavioural classes


Cross-section classification Compression resistance Bending resistance

Class 1

Class 2

Class 3

Class 4

where Wpl is the plastic section modulus, Wel is the elastic section modulus and Weff is the elastic
modulus of an effective cross-section.

Limiting width-to-thickness ratios, to be compared to the actual width-to-thickness ratios of the cross-
section under consideration, are given in Table 5.2 (sheets 1 to 3).
Table 5.2 (sheet 1 of 3): Maximum width-to-thickness ratios for compression parts
Table 5.2 (sheet 2 of 3): Maximum width-to-thickness ratios for compression parts
Table 5.2 (sheet 3 of 3): Maximum width-to-thickness ratios for compression parts

Summary

• Structural sections may be considered as an assembly of individual plate elements.


• Plate elements (parts) may be internal or outstands.
• When loaded in compression these plates may buckle locally.
• Local buckling within the cross-section may limit the load carrying capacity of the section by
preventing the attainment of yield strength.
• Premature failure (arising from the effects of local buckling) may be avoided by limiting the
width-to-thickness ratio of individual elements within the cross-section. This forms the basis of
the cross-section classification approach.
• EC3 defines four classes of cross-section.
• The class into which a particular cross-section falls depends upon the local slenderness of each
individual parts and the compressive stress distribution.
6. COMPRESSION MEMBERS

Background

Common types of compression member cross-section are shown in Figure 6.1.

Figure 6.1: Typical column cross-sections

The design of members for axial compression is more complicated than design for tension due to the
possibility of overall buckling out of the axis of load application, as well as local buckling.
Theoretically, overall flexural buckling would occur in a perfectly straight pin-ended column at the
‘Euler Buckling Load’, as shown by the dashed line in Figure 6.2. However, if the member is of low
slenderness (i.e. short and stocky), then the perfect column would squash at the yield stress rather
than buckle, as shown by the horizontal dashed line in Figure 6.2. From the above discussion and
with reference to Figure 6.2:
Figure 6.2: Failure of perfect pin-ended columns

The axial compression resistance of stocky columns (i.e. low λ ), which corresponds to material
yielding (i.e. the squash load), is given by Equation 6.1.

Nc,Rd = A fy (6.1)

where A is the cross-sectional area (= Aeff for Class 4 sections to account for local buckling) and fy is
the material yield strength.

Figure 6.3 illustrates the load-displacement response for stocky and slender columns.

Figure 6.4: Load-displacement response of columns

The axial compression resistance of slender ‘perfect’ columns (i.e. high λ ), which corresponds to
the Euler critical buckling load, is given by Equation 6.2.

π 2 EI
Ncr = (6.2)
L2cr
where E is the material Young’s modulus, I is the second moment of area about the axis under
consideration (generally the weaker axis), and Lcr is the buckling length of the column.

The non-dimensional member slenderness λ is defined by Equation 6.3.

Af y
λ = (6.3)
N cr

However, the actual strength of columns does not exactly follow the theory (i.e. material yielding or
Euler critical buckling) due to several factors, listed below. Thus, design curves, as shown in Figure
6.3 and explained in the next section, lie on or below the theoretical curves.

A comparison of test results with the theory is shown in Figure 6.4.


Figure 6.4: Form of strength curves for struts with initial out-of-straightness, eccentricity, strain
hardening and residual stresses

Residual stresses are induced primarily during the production process. For cold-formed sections,
residual stresses are principally induced through plastic deformation whereas for hot-rolled and
welded sections, uneven cooling is the main source of residual stresses. Shrinkage of the late-cooling
regions induces residual compressive stresses in the early cooling regions which are then
equilibrated by tensile stresses in the late-cooling regions.

It should be recalled that whether a column is short or long (i.e. low λ or high λ ), if it has a Class 4
cross-section, it may be susceptible to local buckling and therefore require the calculation of an
effective section. Susceptibility of a cross-section to local buckling was discussed last week and
depends on the width/thickness ratios of the individual cross-sectional plate components. It should
also be noted that torsional and flexural-torsional buckling may occur in some members with
inherently low torsional stiffness, such as angle and channel sections.
Design to Eurocode 3

To verify the adequacy of compression members, two types of checks are required:

• Cross-section resistance
• Member buckling resistance

For short compression members ( λ ≤ 0.2), only the cross-sectional check is required, but for longer
members, both checks should be carried out.

Cross-section resistance

Cross-section resistance in compression is covered in clause 6.2.4 of EN 1993-1-1. This of course


ignores overall member buckling effects and therefore may only be applied as the sole check to
members of low slenderness ( λ ≤ 0.2). For all other cases, checks also need to be made for
member buckling as defined in clause 6.3.

Design compressive force is denoted by NEd (axial design effect). The design resistance of a cross-
section under uniform compression, Nc,Rd is determined in a similar manner to BS 5950 Part 1. The
Eurocode 3 design expressions for cross-section resistance under uniform compression are as
follows:

Afy
Nc,Rd = for Class 1, 2 or 3 cross-sections (6.4)
γ M0

A eff f y
Nc,Rd = for Class 4 cross-sections (6.5)
γ M0

For Class 1, 2 and 3 cross-sections, the design compression resistance is taken as the gross cross-
sectional area multiplied by the nominal material yield strength and divided by the partial factor γM0,
and likewise for Class 4 cross-sections with the exception that effective section properties are used
in place of the gross properties.

Member buckling resistance

To verify member buckling resistance, NEd (the axial design effect) must be shown to be less than or
equal to the design buckling resistance of the compression member, Nb,Rd (axial buckling resistance).
Members with non-symmetric Class 4 cross-sections have to be designed for combined bending and
axial compression because of the additional bending moments, ΔMEd that result from the shift in
neutral axis from the gross cross-section to the effective cross-section (multiplied by the applied
compression force).
Compression members with Class 1, 2 and 3 cross-sections and symmetrical Class 4 cross-sections
follow the provisions of clause 6.3.1, and the design buckling resistance should be taken as:

χAfy
N b,Rd = for Class 1, 2 and 3 cross-sections (6.6)
γ M1

χ A eff f y
N b ,Rd = for (symmetric) Class 4 cross-sections (6.7)
γ M1

where χ is the buckling reduction factor . The partial factor γM1 may be taken as 1.0.

Buckling curves

The buckling curves defined by Eurocode 3 Part 1.1 are equivalent to those set out in BS 5950: Part 1
(2000) in tabular form in Table 24 (with the exception of buckling curve a0 which does not appear in
BS 5950). The basic formulations for the buckling curves are as given below:

1
χ = but χ ≤ 1.0 (6.8)
Φ + Φ 2 - λ2

where:

Φ = 0.5[ 1 + α ( λ − 0.2) + λ2 ]

Afy
λ = for Class 1, 2 and 3 cross-sections
N cr

A eff f y
λ = for Class 4 cross-sections
N cr

α is an imperfection factor

Ncr is the elastic critical buckling force for the relevant buckling mode based on the
gross properties of the cross-section.

As shown in Figure 6.5, EN 1993-1-1 defines 5 buckling curves, labelled a0, a, b, c and d. The shapes
of these buckling curves are altered through the imperfection factor α; the 5 values of the
imperfection factor α for each of these curves are given in Table 6.1. Imperfection factors α are
assigned on the basis of cross-section type, axis of buckling and material properties, and may be
found from Table 6.2.
Figure 6.5: Eurocode 3 Part 1.1 buckling curves

From the shape of the buckling curves given in Figure 6.5 it can be seen, in all cases, that for values
of non-dimensional slenderness λ ≤ 0.2, the buckling reduction factor is equal to unity. This means
that for compression members of stocky proportions ( λ ≤ 0.2) there is no reduction to the basic
cross-section resistance. In this case, buckling effects may be ignored and only cross-sectional
checks need be applied.

Table 6.1: Imperfection factors for buckling curves (Table 6.1 of EN 1993-1-1)
Buckling curve a0 a b c d

Imperfection factor α 0.13 0.21 0.34 0.49 0.76

The choice as to which buckling curve (imperfection factor) to adopt is dependent upon the
geometry and material properties of the cross-section and upon the axis of buckling. The
appropriate buckling curve should be determined from Table 6.2 (Table 6.2 of EN 1993-1-1), which is
equivalent to the ‘allocation of strut curve’ table (Table 23) of BS 5950: Part 1 (2000).
Table 6.2: Selection of buckling curve for a cross-section (Table 6.2 of EN 1993-1-1)
Limits Buckling curve
Buckling
S 235
Cross-section about
S 275 S 460
axis
S 355
S 420
b y–y a a0
z tf ≤ 40 mm
z-z b a0

h/b > 1.2


y–y b a
Rolled I-sections

tw 40 mm < tf ≤ 100 mm
z-z c a
h y y
y–y b a
tf ≤ 100 mm
z-z c a

h/b ≤ 1.2
r tf y–y d c
tf > 100 mm
z z-z d c

z z
y–y b b
tf ≤ 40 mm
z-z c c
Welded I-
sections

y y y y
y–y c c
tf tf tf > 40 mm
z-z d d
z z
Hollow sections

hot finished any a a0

cold formed any c c

z
tf generally (except as below) any b b
Welded box
sections

h y y
thick welds: a > 0.5tf
tw
b/tf < 30 any c c
b z h/tw < 30
U-, T- and solid
sections

any c c
L-sections

any b b

Buckling lengths Lcr

The above discussion on column behaviour relates to pin-ended columns. However, in practice, a
variety of other end conditions are possible. End conditions are taken into account in design through
the ‘effective length’ or ‘buckling length’ concept. Typical effective (buckling) lengths for columns with
different end restraint are provided in Table 6.3, where L is the actual (system) length of the column.
Boundary conditions and corresponding buckling lengths are illustrated in Figure 6.6.

Table 6.2: Nominal buckling lengths Lcr for compression members


End restraint (in the plane under consideration) Buckling length Lcr

Effectively restrained in direction at both ends 0.7 L


Effectively
Partially restrained in direction at both ends 0.85 L
held in
position at
Restrained in direction at one end 0.85 L
both ends
Not restrained in direction at either end 1.0 L

One end Other end

Effectively Effectively restrained in direction 1.2 L


held in Not held
position and in Partially restrained in direction 1.5 L
restrained in position
Not restrained in direction 2.0 L
direction

Figure 6.6: Nominal buckling lengths Lcr for compression members


Column design procedure

The following steps may be used for designing a column:

Example/Tutorial 6.1 - Compression member design


7. BEAMS

General

Beams are the most common type of structural member and are intended to span across two (or
more) supports and to transmit the loads mainly by bending action. Table 7.1 summarises the main
types of steel beam indicating the range of spans for which each is most appropriate.

Table 7.1: Typical span and application of various beam types


Beam Type Span Range Notes
(m)
Angles 3–6 Used for roof purlins, sheeting rails etc., where only light load
have to be carried.
Cold-formed 4–8 Used for roof purlins, sheeting rails etc., where only light load
sections have to be carried.
Hot-rolled section Most frequently used type of sections, with proportions selected
1 – 30
to eliminate several possible types of failure.
UBs and UCs
Castellated and Used for long spans and/or light loads. Depth of UB increased by
6 – 60
50% and the web openings may be used for services etc.
cellular beams
Compound 5 –15 Used when a single rolled section would not provide sufficient
sections (e.g. UB bending strength (capacity).
Made by welding together plates (usually 3). Web depths up to
Plate girders 10 – 100
3-4 m; webs sometimes need stiffening.

Made up of chords and diagonals fabricated from tubes, angles


Trusses 10 - 100
or hot-rolled sections. The chords and diagonals are subjected
principally to axial forces.
Fabricated from plates and usually stiffened. Used for crane rails
Box girders 15 – 200 and bridge decks due to high torsional and transverse stiffness
properties.

As with compression members, a crucial parameter to define the behaviour of flexural (beam)
members is their non-dimensional slenderness. For columns, the non-dimensional slenderness is
denoted by λ , whereas for beams λ LT is adopted to reflect the mode of buckling to which beams
are generally susceptible (Lateral Torsional Buckling). Analogous to columns, beam members with
low non-dimensional slenderness (i.e. low λ LT ) will failure by material yielding (which, for the case
of beams corresponds to in-plane failure), whilst beams with high non-dimensional slenderness (i.e.
high λ LT ) will fail by member buckling (which, for the case of beams corresponds to lateral torsional
buckling). In plane failure is characterised by excessive bending and deformation in the plane of the
applied loading

Lateral torsional buckling is characterised by twisting and lateral deflection of a beam that is
subjected to bending about its major axis (See Figure 7.1). Lateral torsional buckling may be
prevented by providing full lateral restraint to the beams. For laterally unrestrained beams, lateral
torsional buckling must be checked and designed for.

Figure 7.1: Lateral torsional buckling of a slender cantilever beam

Full lateral restraint is commonly achieved by the provision of sufficient lateral bracing to the
compression flange of the beam or where the beam supports a concrete slab. There are a number
of other situations where lateral torsional buckling need not be considered – a full list is given below:

• where sufficient lateral restraint is applied to the compression flange of the beam
• where bending is about the minor axis
• where cross-sections with high lateral and torsional stiffness are employed, for example square
or circular hollow sections
• or generally where the non-dimensional lateral torsional slenderness, λ LT ≤ 0.2 (or in some
cases λ LT ≤ 0.4 (see clause 6.3.2.3 of EN 1993-1-1)).

In cases where LTB is a problem, bracing may be used as a means of improving performance and
providing lateral stability. Two requirements are necessary:

1. The bracing must be sufficiently stiff to hold the braced point effectively against lateral
movement (this can normally be achieved without difficulty).
2. The bracing must be sufficiently strong to withstand the forces transmitted to it by the main
member (these forces are normally a percentage of the force in the compression flange of the
braced beam member).

For cases where lateral torsional buckling does not occur, member bending strength (resistance)
may be assessed on the basis of the in-plane cross-sectional resistance alone, for which, as described
later, the cross-section classification is important in defining the beams behaviour.

As well as bending moment resistance, there are two other common checks that should be carried
out to verify the suitability of a beam to support applied loading, as given below.

• Bending moment resistance


• Shear resistance
• Deflections.

In addition, the resistance of the beam to lateral torsional buckling may need to be checked.

Design to Eurocode 3

In-plane bending of beams

For beams with stocky cross-sections, the load-deflection behaviour will be similar to that shown in
Figure 7.2. The beam response will be linear until the maximum extreme fibre stress at the critical
cross-section reaches yield. At higher loads, the deformations will increase more rapidly until the fully
plastic moment Mpl is reached at the critical section forming a plastic hinge.
Figure 7.2: Load- deflection behaviour of a simply supported beam

However, a cross-section will only be able to reach its plastic moment resistance if it possesses
sufficient rotation (deformation) capacity. Whether a cross-section has sufficient rotation capacity is
assessed through its cross-section classification. Cross-section classification was discussed in detail
in Section 5 of the notes, and the key points are repeated below:

• Class 1 cross-sections are fully effective under pure compression and are capable of reaching
and maintaining their full plastic moment Mpl in bending (and may therefore be used in plastic
design).
• Class 2 cross-sections have a somewhat lower deformation capacity, but are also fully effective
in pure compression and are capable of reaching their full plastic moment Mpl in bending.
• Class 3 cross-sections are fully effective in pure compression, but local buckling prevents
attainment of the full plastic moment in bending; bending moment resistance is therefore
limited to the (elastic) yield moment Mel.
• For Class 4 cross-sections, local buckling occurs in the elastic range. An effective cross-section is
therefore defined based on the width-to-thickness ratios of individual plate elements, and this is
used to determine the cross-sectional resistance Meff.

In hot-rolled design the majority of standard cross-sections will be Class 1, 2 or 3, where resistances
may be based on gross section properties obtained from section tables (see Appendix A of these
notes). Effective width formulations are not contained in EN 1993-1-1 (Part 1.1 of Eurocode 3), but
may instead be found in EN 1993-1-5 (Part 1.5 of Eurocode 3).

The distributions of elastic and plastic bending stresses for symmetric and asymmetric sections are
given in Figures 7.3 and 7.4. The cross-section of the beam should be chosen such that the bending
moment resistance (Mc,Rd) is higher than the applied design moment (MEd).
Figure 7.3: Distribution of elastic and plastic bending stresses about an axis of symmetry

Figure 7.4: Distribution of elastic and plastic bending stresses about an axis of asymmetry

Eurocode 3 adopts the symbol W for all section moduli. Subscripts are used to differentiate between
the plastic, elastic or effective section modulus (Wpl, Wel, or Weff respectively). The partial factor γM0
is applied to all cross-section bending resistances. As in BS 5950 Part 1, the resistance of Class 1 and
2 cross-sections is based upon the full plastic section modulus, the resistance of Class 3 cross-sections
is based upon the elastic section modulus, and the resistance of Class 4 cross-sections utilises the
effective section modulus.
The design expressions are given below:

Wpl f y
Mc,Rd = for Class ?? cross-sections (7.1)
γ M0

Wel f y
Mc,Rd = for Class ? cross-sections (7.2)
γ M0

Weff f y
Mc,Rd = for Class ? cross-sections (7.3)
γ M0

where the section moduli W should be obtained from section tables, the material yield strength fy
should be taken from Table 1.1 of the notes (depending upon the grade of the steel and the material
thickness) and γM0 is equal to 1.0.

Lateral torsional buckling (LTB)

Lateral torsional buckling is the member buckling mode associated with slender beams loaded about
their major axis, without continuous lateral restraint. As noted above, if continuous lateral restraint
is provided to the compression flange of the beam, then lateral torsional buckling will be prevented
and failure will occur in another mode, often in-plane bending (and/or shear). The Eurocode design
method accounts for the large number of parameters that can influence LTB including section shape,
beam geometry, the degree of lateral restraint, support conditions, the type of loading (normal or
destabilising) as well as the residual stress pattern and initial imperfections.

Lateral torsional buckling curves for the general case are described through Equation 7.4. The
general design case applies to all section types, while the ‘case for rolled and equivalent welded
sections’, which provides slightly modified buckling curves, can be used for hot-rolled I, H and hollow
sections and welded sections of similar proportions. The basic design approach is the same for both
cases. THe design buckling resistance moment of a laterally unrestrained beam should be taken as:

fy
M c , Rd = χ LT W y (7.4)
γ M1

where Wy is the appropriate section modulus as:

Wy = Wpl,y for Class 1 or 2 sections

Wy = Wel,y for Class 3 sections

Wy = Weff,y for Class 4 sections

χLT is the reduction factor for lateral torsional buckling, given as:
1
χ LT = but χLT ≤ 1.0 (7.5)
Φ LT + Φ 2LT − λ2LT

where: ΦLT = 0.5 [ 1 + α LT ( λ LT − 0.2) + λ2LT ]

Wy f y
λ LT =
M cr

in which αLT is an imperfection factor from Table 7.2 (Table 6.3 of EN 1993-1-1) and Mcr is the elastic
critical moment for lateral torsional buckling (see the following sub-section).

The imperfection factors αLT for the 4 lateral torsional buckling curves are given by Table 7.2 (Table
6.3 of EN 1993-1-1), and choice of buckling curves should be made with reference to Table 7.3.

Table 7.2: Imperfection factors for lateral torsional buckling curves (Table 6.3 of EN 1993-1-1)
Buckling curve a b c d

Imperfection factor αLT 0.21 0.34 0.49 0.76

Table 7.3: Lateral torsional buckling curve for cross-sections (Table 6.4 of EN 1993-1-1)
Cross-section Limits Buckling curve

Rolled I-sections h/b ≤ 2 a


h/b > 2 b
Welded I-sections h/b ≤ 2 c
h/b > 2 d
Other cross-sections - d

Elastic critical moment for lateral torsional buckling Mcr


As shown in the previous section, determination of the non-dimensional lateral torsional buckling
slenderness λ LT first requires calculation of the elastic critical moment for lateral torsional buckling
Mcr. However, Eurocode 3 offers no formulations and gives no guidance on how Mcr should be
calculated, except to say that Mcr should be based on gross cross-sectional properties and should
take into account the loading conditions, the real moment distribution and the lateral restraints
(clause 6.3.2.2(2)). Reasons for the omission of such formulations include the complexity of the
subject and a lack of consensus between the contributing nations; by many, it is regarded as
‘textbook material’.

The elastic critical moment for lateral torsional buckling of a beam of uniform symmetrical cross-
section with equal flanges, under standard conditions of restraint at each end, loaded through the
shear centre and subject to uniform moment is given by equation 7.6. The standard conditions of
restraint at each end are: restrained against lateral movement, restrained against rotation about the
longitudinal axis and free to rotate on plan. Equation 7.6 was provided in ENV 1993-1-1 (1992) in an
Informative Annex, and has been shown, for example by Timoshenko and Gere, to represent the
exact analytical solution to the governing differential equation.

0.5
π 2 EI z  I w L cr GI T 
2
Mcr,0 =  +  (7.6)
L cr 2  I z π 2 EI z 

E
where: G =
2(1 + υ)
IT is the torsion constant
Iw is the warping constant
Iz is the second moment of area about the minor axis
Lcr is the length of the beam between points of lateral restraint

Numerical solutions have also been calculated for a number of other loading conditions. For
uniform doubly-symmetric cross-sections, loaded through the shear centre at the level of the
centroidal axis, and with the standard conditions of restraint described above, Mcr may be calculated
through Equation 7.7.

0.5
π 2 EI z  I w L cr GI T 
2
Mcr = C1  +  (7.7)
L cr 2  I z π 2 EI z 

where C1 may be determined from Table 7.4 for end moment loading and from Table 7.5 for
transverse loading. The C1 factor is used to modify Mcr,0 (i.e. Mcr = C1 Mcr,0) to take account of the
shape of the bending moment diagram, and performs a similar function to the ‘m’ factor adopted in
BS 5950. Other extensions of Equations 7.6 and 7.7 are available to account for different loading
positions, level of the load applications etc. but these are not included in this course (see
recommended text 2 for more information).

The values of C1 given Table 7.4 for end moment loading may be approximated by equation 7.8,
though other approximations also exist.

C1 = 1.88 – 1.40ψ + 0.52ψ 2 but C1 ≤2.70 (7.8)

where ψ is the ratio of the end moments (defined in Table 7.4).

Figure 7.5 compares values of C1 obtained from Table 7.4 and from Equation 7.8. Figure 7.5 shows,
as expected, that the most severe loading condition (that of uniform bending moment where
ψ = 1.0, results in the lowest value for Mcr. As the ratio of the end moments ψ decreases, so the
value of Mcr rises; these increases in Mcr are associated principally with changes that occur in the
buckled deflected shape, which change from a symmetric half sine wave for uniform bending
moment (ψ = 1) to an anti-symmetric double half wave for ψ = -1. At high values of C1, there is some
deviation between the approximate expression (equation 7.8) and more the accurate tabulated
results of Table 7.4; thus Equation 7.8 should not be applied when C1 is greater than 2.70.

Figure 7.5: Tabulated and approximate values of C1 for varying ψ


Table 7.4: C1 values for end moment loading
Loading and support Bending moment diagram Value of C1
conditions
ψ=+1
1.000

ψ = + 0.75
1.141

ψ = + 0.5
1.323

ψ = + 0.25
1.563

ψ=0
1.879
M ψM

ψ = - 0.25
2.281

ψ = - 0.5
2.704

ψ = - 0.75
2.927

ψ=-1
2.752
Table 7.5: C1 values for transverse loading
Loading and support Bending moment diagram Value of C1
conditions

w
1.132

w 1.285

F
1.365

F
1.565

CL
F F
1.046
= = = =

Shear

Except in cases of high coincident shear and moment, such as at the internal supports of continuous
beams, the effect of shear is unlikely to have a significant influence on the design of beams.
However, it should always be checked.

Actual (elastic) shear stress distributions developed in rectangular sections and I-sections are as
shown in Figure 7.6. In both cases shear stress varies parabolically with depth, with maximum value
occurring at the neutral axis. For the I-section, as with most common structural steel sections, the
difference between maximum and minimum values for the web, which carries almost all of the
vertical shear force, is sufficiently small for design to be simplified by working with the average shear
stress (and assuming a degree of plastic redistribution). Average shear stress is defined as the total
shear force VEd divided by the area of the web (or equivalent shear area Av).
Figure 7.6: Distribution of shear stresses in beams subjected to a shear force VEd

Since the yield stress of steel in shear is approximately 1/√3 of its yield stress in tension, Eurocode 3
(clause 6.2.6) therefore defines the plastic shear resistance as:

A v (f y 3)
Vpl,Rd = (7.9)
γ M0

and it is the plastic shear resistance that will be used in the vast majority of practical design
situations.

The shear area Av is in effect the area of the cross-section that can be mobilised to resist the applied
shear force with a moderate allowance for plastic redistribution, and for sections where the load is
applied parallel to the web, this is essentially the area of the web (with some allowance for the root
radii in rolled sections). Expressions for the determination of shear area Av for general structural
cross-sections are given in clause 6.2.6(3) of EN 1993-1-1. The most common ones are repeated
below:
Rolled I and H sections, load parallel to web: Av = A – 2btf + (tw + 2r)tf but ≥ ηhwtw

Rolled channel sections, load parallel to web: Av = A – 2btf + (tw + r)tf

Welded I, H and box sections, load parallel to web: Av = ηΣ(hwtw)

Welded I, H, channel and box sections, load parallel to flanges: Av = A - Σ (hwtw)

Rolled RHS of uniform thickness, load parallel to depth: Av = Ah/(b+h)

Rolled RHS of uniform thickness, load parallel to width: Av = Ab/(b+h)

CHS and tubes of uniform thickness: Av = 2A/π

where:
A is the cross-sectional are
b is the overall section breadth
h is the overall section depth
hw is the overall web depth (measured between the flanges)
r is the root radius
tf is the flange thickness
tw is the web thickness
η may conservatively be taken as 1.0 (or 1.2 according to EN 1993-1-5)

The presence of shear in a section theoretically reduces its moment capacity. In practice, however,
this reduction may be neglected up to large fractions of the shear resistance Vpl,Rd. Clause 6.2.8(2) of
EN 1993-1-1 states that provided the applied shear force is less than half the plastic shear resistance
of the cross-section its effect on the moment resistance may be neglected.

Deflections

Excessive serviceability deflections may impair the function of a structure, for example, leading to
cracking of plaster, misalignments of crane rails, causing difficulty in opening doors, etc. Deflection
checks should therefore be performed against suitable limiting values. Typical vertical and
horizontal and deflection limits are given in Tables 7.6 and 7.7, respectively. Some degree of
approximation in calculating deflections is usually acceptable, but checks should be made against
unfactored variable actions Qk.

Table 7.6: Vertical deflection limits


Design situation Deflection limit
Cantilevers Length/180
Beams carrying plaster or other brittle finish Span/360
Other beams (except purlins and sheeting rails) Span/200
Purlins and sheeting rails To suit cladding
Table 7.7: Horizontal deflection limits
Design situation Deflection limit
Tops of columns in single storey buildings, except Height/300
portal frames
Columns in portal frame buildings, not supporting To suit cladding
crane runways
In each storey of a building with more than one Height of storey/300

Formulae for determining the deflections of beams with common loading and support arrangements
are given by Table 7.8.

Table 7.8: Formulae for determining beam deflections for standard cases

Beam and loading Deflections

Midspan deflection:
w

5 wL4
δ =
384 EI
L

F Midspan deflection:

FL3
δ =
48 EI
L/2 L/2

Deflection at cantilever tip:


w
wL4
δ =
8 EI
L

Deflection at cantilever tip:


F
FL3
δ =
3 EI
L

Example 7.1 - Beam design


Example 7.1 demonstrates the calculations required (to Eurocode 3) to check the resistance beam
undergoing bending.
Plate girders

Plate girders are structural sections, fabricated (welded) from (usually 3) steel plates, and employed
to support heavy loads over long spans (see Figure 7.7). They are generally adopted when existing
hot-rolled structural members have insufficient moment resistance or flexural stiffness to support
the applied loads. Typical applications include transfer beams in buildings and small to medium
span bridges.

Figure 7.7: Plate girder being fabricated (photo courtesy of Allerton Steel Ltd)

The overall proportions of the plate girder and the proportions of the individual plate elements are
the choice of the designer, with the general aim of achieving minimum material cost, though other
restraints (e.g. height restrictions) may also exist. For efficiency, thick flanges, separated by a deep,
thin web is ideal, but careful consideration of buckling effects, particularly shear buckling of the web,
need to be made.

The following rules of thumb may be used for initial sizing purposes:

• For simply-supported plate girders, a span-to-depth ratio of between about 15 and 20 will
generally provide economic solutions. For particularly high loading (for example in railway
bridges) a lower value may be suitable.

• Initial sizing of the flanges may be achieved by ignoring the contribution to bending resistance of
the beam web, and calculating the moment resistance of the flanges alone (i.e. flange area ×
yield strength × lever arm between flanges). For initial sizing, and since the flange proportions
are, as yet, not established, the lever arm may be taken as the web depth hw.
M max
Therefore, for a single flange, Af ≈
hw fy

where: Mmax is the maximum bending moment along the length of the beam (i.e. MEd)

hw is the depth of the web between the flanges

fy is the material strength

• Having established the required area of the flanges Af, the flanges may now be proportioned. It
is recommended that the proportions are chosen such that their classification is Class 1 or 2 (or
Class 3) – see Table 5.2 of notes, but Class 4 is best avoided to ensure that effective section
calculations are not necessary. If Class 1 or 2 proportions are selected, and the web (in bending)
is also Class 1 or 2, then the moment resistance may be based on the full plastic bending
resistance. For Class 3 cross-sections, the elastic bending strength should be adopted.

• For maximum efficiency (in terms of minimisation of material) plate girders webs are generally
very slender and often with transverse (vertical web) stiffeners at regular intervals; shear
buckling resistance is then checked against the maximum design shear force. However, to
simplify design calculations, shear buckling may be ignored provided the web proportions satisfy
the limitations below:

hw ε
≤ 72
tw η
where: hw is the depth of the web between the flanges

tw is the web thickness

ε = 235 / f y

η = 1.2 according to EN 1993-1-5, or may conservatively be taken as 1.0

A minimum web thickness of 6 mm is recommended.

Final checks to demonstrate that the bending resistance, shear resistance and deflection
requirements are met, should be made as usual. All required section properties, such as plastic
modulus Wpl and second moment of area Iy need to be determined by calculation.
Summary of beams

Beams may be considered as restrained if:

• sufficient lateral restraint is applied to the compression flange of the beam, for example by a
floor system or a profiled roof sheeting;
• bending is about the minor axis of the beam (which cannot fail by lateral-torsional instability);
• cross-sections with high lateral and torsional stiffness are employed, for example square or
circular hollow sections;
• generally, the non-dimensional lateral torsional slenderness, is less than 0.2.

The procedure for designing laterally restrained beams is as follows:

• Check for cross-section resistance in bending;


• Check for cross-section resistance in shear;
• Check for the limit against shear buckling resistance;
• Check for cross-section resistance in combined bending and shear.

The procedure for designing unrestrained beams is identical to that for restrained beams except that
one additional check is required, to ensure that the lateral torsional buckling resistance is not
surpassed.

Example 7.2 – LTB resistance design


Example 7.2 demonstrates the check for lateral torsional buckling in a simply supported beam.
Tutorial 7.1 – Load distribution and beam design

For the single storey building shown in Tutorial 6.2 (and repeated below), it is required to design
beams B2 and B4 for the applied loads, making all necessary code checks (i.e. section classification,
bending, shear and deflections). The permanent (dead) load from the roof slab is 2.5 kN/m2 and the
variable imposed load is 1.5 kN/m2. For the unfactored self-weight of the beams, assume 0.4 kN/m
for the secondary beams, and 0.8 kN/m for the primary beams. Use a deflection limit of span/200.
8. JOINTS – BOLTED CONNECTIONS

Joints are necessary to form structural continuity from a discontinuous set of members or
components, as well as to overcome size limitations in manufacture and transportation. Joints may
be formed by bolting or welding or a combination of both. This section deals with bolted
connections while the following section describes welded connections.

Introduction

This course provides an overview of the behaviour and design of bolted and welded (next Section)
connections. Eurocode 3 Part 1-8 covers the design of joints. The following are relevant definitions
from that code:

Connection (Clause 1.4.2): Location at which two or more elements meet. For design purposes it is
the assembly of the basic components required to represent the behaviour during the transfer of
the relevant internal forces and moments at the connection.

Joint (Clause 1.4.4): Zone where two or more members are interconnected. For design purposes it is
the assembly of all the basic components required to represent the behaviour during the transfer of
the relevant internal forces and moments between the connected members.

There are several standard connection details and some examples are shown in Figure 8.1.

(a)

(b)

Figure 8.1: Examples of standard joints (a) beam-column joints (b) bracing joints
Influence of joint type on structural behaviour

The stiffness and strength of the joints have an important influence on the behaviour of structures.
Three different types of joints are recognised in Eurocode 3, (pinned, semi-rigid and rigid) and
examples of each are shown in Figure 8.2(a) while Figure 8.2(b) illustrates the moment-curvature
relationships.

(a)

(b)

Figure 8.2: (a) different types of joint and (b) moment-rotation characteristics of joints

Joints defined as ‘nominally pinned’ are incapable of transmitting significant moment, but are able
to accept the resulting rotations under design loads. ‘Rigid and full strength’ joints have sufficient
stiffness to assume full continuity between members, and strength at least equal to the strength of
the connected members. ‘Semi-rigid and partial strength’ joints are intermediate between the two
extremes. The Eurocode directly links the three types of joint with three analysis types (i.e. simple,
semi-continuous and continuous analysis). It does this by classifying the joint types in terms of their
strength (moment resistance) and stiffness.
Types of bolt

Bearing bolts

As shown in Figure 8.3, in a shear connection, load is transferred by bearing action through: (1)
Bearing in plate, (2) Bearing in bolt and (3) Shearing in bolt. For bolts in shear we are interested in
the cross-sectional area of the bolt, whereas in bearing we have to consider the area derived from
the bolt diameter multiplied by the thickness of the connected part.

In a tension connection there is bearing under the bolt head and under the nut, hence load transfer
is by: (1) Bearing under bolt head, (2) Tension in bolt and (3) Bearing under nut.

Figure 8.3: Load transfer in bearing bolts

High Strength Friction Grip bolts (HSFG)

As shown in Figure 8.4, for this type, the bolts are tightened up to their proof load using one of
several approved methods. This clamps the plies (plates) together so that a circular ring of high
friction capacity develops around the bolt. This friction transfers the forces P/2. It is necessary,
however, also to check the bearing capacity of the plate in case the bolt slips into contact with the
side of the hole.

In the tension connection, HSFG is advantageous if a cyclic load is applied (i.e. due to traffic, wind,
machinery vibration, etc.). A bearing bolt with cyclic tension would have low fatigue life because of
the stress concentration in the threads. The fatigue life is a function of the magnitude of the cyclic
component of tension in the bolt not of the absolute magnitude of tension in the bolt. Therefore,
using HSFG will reduce the cyclic variation of the bolt force and hence considerably enhance its
fatigue life.

Figure 8.4: Load transfer in HSFG bolts

Bolt grades and sizes

Bearing bolts

The most common classes of bolt are 4.6 (black bolts) and 8.8 (high strength bolts). The yield and
ultimate strength of bolts may be determined directly from their class, since each class is described
by two numbers separated by a point X.Y, with X representing the ultimate tensile strength fub and .Y
is the yield strength fyb, expressed as a fraction of the ultimate tensile strength). For example, a class
8.8 bolt has an ultimate tensile strength fub of 800 N/mm2 and a yield strength fyb of 640 N/mm2).
Bolt size is given as the diameter of the shank (the unthreaded portion of the bolt) in mm, for
example M12, M16, M20, M24, M30, M36, with the ‘M’ designating ‘metric’.
High Strength Friction Grip (HSFG) bolts

High strength friction grip bolts use the same classification system as bearing bolts, but are generally
of higher strength. For HSFG bolts, the ‘general grade’ is made of equivalent material to 8.8, but
higher grades are not uncommon (for example 10.9).

Geometric Considerations

Generally ‘clearance’ bolt holes are adopted, whereby the bolt hole is marginally larger than the
nominal bolt diameter, as given below. In some circumstances, oversize or slotted holes will be
specified.

• Clearance of 2 mm for bolt diameter ≤


• Clearance of 3 mm for bolt diameter >
• Oversize or slotted holes may be used in some cases, but will affect slip resistance.

Spacing and edge and end distances

During load transfer, the bolt bears into the plate causing the bolt hole to elongate in the direction
of the load. If the bolt hole is adjacent to the edge of the plate or if the spacing between the bolts is
relatively small, the bolt may tear through the plate (see Figure 8.5). To prevent this from occurring
too easily, Eurocode 3 Part 1.8 (EN 1993-1-8) requires minimum edge distance and spacing between
bolts.

Figure 8.5: Bolt tearing through plate due to small end distance

Minimum requirements
Minimum bolt spacings and edge and end distances are as below, where d0 is the fastener (bolt) hole
diameter (see Figure 8.6). These values are defined in Table 3.3 of EN 1993-1-8.

• Minimum spacing of bolts in the direction of load transfer p1 = 2.2d0


• Minimum end distance in the direction of load transfer e1 = 1.2d0
• Minimum spacing of bolts perpendicular to the direction of load transfer p2 = 2.4d0
• Minimum edge distance perpendicular to the direction of load transfer e2 = 1.2d0

Figure 8.6: Definitions for p1, e1, p2 and e2

Maximum spacing and edge distances are also specified in Eurocode 3 to ensure appropriate load
and stress distribution.

Backmarks (distance from back of angle or channel web to the centre of a hole through the leg or
flange) are usually specified to be close to the centroidal axis. Also, cross centres (distances between
hole centres in flanges of UBs and UCs) are specified considering accessibility and edge distances.
Examples are shown in Figures 8.7 and 8.8.

Design of bolted connections to Eurocode 3

For shear connections, the following items should be checked:

• Minimum (and maximum) bolt spacing and edge and end distances
• Shear resistance Fv,Rd (or slip resistance Fs,Rd) of the bolts
• Bearing resistance Fb,Rd
• Resistance of the member (at gross and net cross-sections)

Other checks, such as for combined loading, block tearing etc. may also be necessary.

For tension connections, in addition to the minimum (and maximum) bolt spacing and edge and end
distances, the tensile resistance of the bolts should be checked, with due allowance made for prying
action.
Shear resistance Fv,Rd and slip resistance Fs,Rd

Shear stresses are developed across the cross-sectional area of the bolt. In Figure 8.8(a) there is only
one shear plane hence all load is carried on it (Single Shear). In Figure 8.8(b) there are two shear
planes where P/2 is carried on each (Double Shear).

Figure 8.8: Bolts loaded in (a) single shear and (b) double shear

Unless it is ensured that the shear plane falls outside the threaded portion then the cross-sectional
area should be assumed based on the threaded area.

Table 3.4 of EN 1993-1-8 defines bolt shear resistance per shear plane for ordinary bolts as:

α v f ub A
Fv,Rd = (8.1)
γ M2

where:

αv = 0.6 for classes 4.6, 5.6 and 8.8 where the shear plane passes through the
threaded portion of the bolt, and for all classes where the shear plane passes
through the unthreaded portion of the bolt
= 0.5 for classes 4.8, 5.8, 6.8 and 10.9 where the shear plane passes through
the threaded portion of the bolt
fub is the ultimate tensile strength of the bolt
A is the shear stress area As (i.e. area at threads – see Table 8.1) when the shear
plane passes through the threaded portion of the bolt or the gross cross-
sectional area when the shear plane passes through the unthreaded (shank)
portion of the bolt.
γM2 may be taken as

For HSFG bolts, shear force is resisted by friction until slip occurs. The slip resistance Fs,Rd is given by:

ks n µ
Fs,Rd = Fp ,Cd (8.2)
γ M3
in which:

n is the number of friction surfaces (i.e. n = 1, when two plates are joined)
Fp,Cd is the minimum pre-load force in the bolt = 0.7 × 800 As
γM3 may be taken at

Values for factor ks as well as a set of the slip factor μ corresponding to four classes of plate surface
are provided in Tables 3.6 and 3.7 of EN 1993-1-8, respectively. For bolts in normal holes, ks = 1.0.
Four classes of friction surfaces, with values of μ ranging from 0.2 to 0.5 are listed in Table 3.7 of EN
1993-1-8. For a shot blasted surface (Class A), a value of μ = 0.5 may be used.

Bearing resistance Fb,Rd

Bearing resistance is governed by the projected contact area between a bolt and connected parts,
the ultimate material strength (of the bolt or the connected parts), and may be limited by bolt
spacing and edge and end distances.

From EN 1993-1-8, bearing resistance is given by:

k 1 α b f u dt
Fb, Rd = (8.3)
γ M2

in which :

αb is the smallest of: αd; fub/fu or 1.0, and accounts for the possibility of
reduction in bearing resistance due to the bolt spacing or edge and end
distances, or due to bearing failure in the bolt (controlled by fub), rather than in
the connected parts (controlled by fu).
d is the bolt diameter
t is the thickness of the connected parts (if parts of different thicknesses are
connected, take the smallest value)
γM2 may be taken as 1.25
fu is the ultimate tensile strength of the connected parts and, (with reference to
Figure 8.5):

In the direction of load transfer:


e1
αd = for end bolts
3d 0
p1
αd = − 0.25 for inner bolts
3d 0
Perpendicular to the direction of load transfer:
e2
k1 is the smaller of: ( 2.8 − 1.7 ) or 2.5 for edge bolts
d0
p2
k1 is the smaller of: ( 1.4 − 1.7 ) or 2.5 for inner bolts
d0

Tension resistance Ft,Rd

Upon tension loading, tensile stresses develop in the stressed length of a bolt (i.e. between nut and
head of the bolt). Tension resistance is given in EN 1993-1-8 as:

k 2 f ub A s
Ft, Rd = (8.4)
γ M2

where:

As is the tensile stress area (i.e. area through threads) of the bolt (see Table 8.1)

k2 = 0.9 (except for countersunk bolts where k2 = 0.63)

fub is the ultimate strength of the bolt (determined from the bolt grade).

Table 8.1: Nominal (unthreaded or shank) and threaded areas of bolts


Nominal bolt diameter d Area at threads As (mm2)
Nominal Area A (mm2)
(mm)
M12 113 84.3
M14 154 115
M16 201 157
M18 254 192
M20 314 245
M22 380 303
M24 452 353
M27 573 459
M30 707 561

In some situations, bolts may experience tension and shear in combination. In general, bolt
capacities would be expected to reduce when high values of shear and tension are coexistent. EN
1993-1-8 provides the following interaction expression to deal with such case:

Fv , Ed Ft , Ed
+ ≤ 1.0 (8.5)
Fv , Rd 1.4 Ft , Rd
Bolt tables

Bolt tables, giving individual resistances for different geometric and loading arrangements, are
available in design manuals for convenient design.

Analysis of bolt groups

Previous sections have established the resistances of individual bolts and connected parts in the
region of the bolt. This section deals with determining the forces applied to the bolts in bolt groups.
This is a function of the type of load applied and of the overall configuration of the connection.

Concentrically-loaded shear joints

In concentric shear connections, the bolts should be arranged in a regular symmetric pattern such
that the load axis passes through the centroid of bolt group. In this case, if the effective line of action
of the load is down an axis of symmetry, the load can be shared equally among all bolts (i.e. F = P/n).
An example of this type is the simple tension splice shown in Figure 8.9. For this example, there are 9
bolts (n = 9) in double shear.

Figure 8.9: Simple tension splice (with bolts loaded in shear)

Example/Tutorial 8.1 – Bolted joint design


Example and Tutorial 8.1 relates to the design of concentrically loaded bolt groups.
Eccentrically-loaded shear joints

In this case the bolts act in shear and bearing as in the concentrically-loaded joints, but now the
shear is applied eccentricity to the bolt group centroid. This imposes not only a translation on the
bolt group but a twisting action as well. Therefore, although the bolt continue to act in shear and
bearing, the magnitude of the bolt loading increases, (and the direction of the bolt loading changes).
The analysis may be carried out by replacing the eccentric shear force by a concentric shear force (R)
through the centroid of the bolt group plus a torsion moment about the centroid (of magnitude
R × c). An example for eccentrically-loaded shear connections is shown in Figure 8.10.

Figure 8.10: Eccentrically-loaded shear connection (showing bolt reactions

Example on analysis of eccentric shear connections


Consider the joint shown in Figure 8.10. The eccentric shear force, R, may be replaced by a concentric
vertical shear force (i.e. acting at the centroid of the bolt group) and a centroidal torsional moment,
(M), such that:

M= (where c = eccentricity of shear force from centroid of bolt group)

The shear force (Q) carried by each bolt due to the (concentric) vertical shear force is:

Q = (where n = number of bolts)

The additional shear force due to the torsional moment acts at right angles to the line joining each
bolt to the bolt group centroid. The magnitude of this force (Fmi) for a particular bolt (i) is
proportional to the distance between the bolt and the centroid of the bolt group. See Figure 8.11.
Figure 8.11: Example of bolt group under eccentric shear force (showing forces acting on bolt)

The shear force Fmi in any bolt, i, due to the moment can be shown (by considering equilibrium and
compatibility) to be:

i=n i=n i=n


Fmi =
M ri
i=n
where ∑ ri2 = ∑ x i2 + ∑y 2
i

∑r
i =1
i
2 i =1 i =1 i =1

For a critical bolt (bolt with maximum distance from centroid, closest to the applied load) (i.e. bolt
number 1), the shear force, Fm, due to the moment is:

M r1
Fm1 = i=n

∑r
i =1
i
2

The horizontal and vertical components of Fm, which enable convenient determination of the
resultant force on the bolt, are Fh and Fv respectively, where:

M y1 M x1
Fh1 = i=n
and Fv1 = i=n
∑ (x i2 + y i2 ) ∑ (x i2 + y i2 )
i =1 i =1

The resultant shear force FR (due to both the concentric vertical shear force and the moment) on the
critical bolt is therefore:

FR = (Q + Fv ) 2 + Fh2
For a safe design this resultant force should be lower than the resistance of the bolt (i.e. least of
shear resistance and bearing resistance). If unsafe, increase area or number of bolts, or both.

Connections subjected to out-of-plane moments

Examples of this type are shown in Figure 8.12. In both cases there is an out-of-plane moment
applied to the joint causing the bottom of the end-plate to press against the column flange and the
top of the plate to separate from the column flange. Hence the top row of bolts is subjected to the
largest tensile forces and is more susceptible to early yielding. Design of this type of connection is
not covered in detail in this course.

Figure 8.12: Out-of-plane moments on connections


9. JOINTS - WELDED CONNECTIONS

Welded connections are produced by the melting of both the weld metal and parent material, such
that the weld and connected parts fuse together upon cooling. Welding should be carried out by
skilled operatives and inspected upon completion. There are two basic types of weld: butt welds
and fillet welds.

Types of weld

Butt welds

Butt welds (see Figure 9.1) are formed within the contact area (the butting face) between the pieces
to be welded. If the whole contact area is welded, the weld is described as ‘full penetration butt
weld’ (FPBW). A properly executed FPBW will develop the full strength of the connected part.

Figure 9.1: Butt weld

If only part of the contact area is welded, we have a ‘partial penetration butt weld’ (PPBW), which, in
general, would have a reduced strength. The strength of the PPBW is based on the effective throat of
the weld and consideration should be given to the eccentricity. Note that forming a FPBW or a PPBW
will usually require some edge preparation of plates so that there is access to create the weld on the
butting face (see Figure 9.2).
Figure 9.2: Basic types of butt weld

Figure 9.3: Edge preparation in butt welds

Fillet welds

Fillet welds are not formed on the butt face but are external to the profile of the connected parts, as
shown in Figure 9.4.

Figure 9.4: Fillet welds


Failure of the weld is assumed on the minimum cross-sectional area of the weld, i.e. on the ‘throat
thickness’, a (see Figure 9.5). The size of the weld is usually specified by the ‘leg length’, s. Note that
an ‘8 mm weld’ means a weld of leg length s = 8 mm.

Figure 9.5: Definitions of weld size (leg length), s and throat thickness, a

Fillet welds finishing at the ends or sides of parts should be returned continuously, full size, around
the corner for a distance of at least twice the leg length s (see Figure 9.6) of the weld, unless this is
impractical due to access restrictions.

Figure 9.6: Side (longitudinal) fillet weld

Geometric considerations

Effective weld length, l

The effective length l of a fillet weld should be taken as the length over which the fillet is full size.
Provided the weld is full size throughout its length, including starts and terminations, the effective
weld length may be taken as the actual weld length. Account may be taken for possible reduction in
weld size at its ends by assuming the effective length as the overall length minus twice the effective
throat thickness a.
The minimum effective length for a load bearing fillet weld is 30 mm or 6 times its throat thickness a,
whichever is larger.

The maximum effective length, before reduction factors due to non-uniformity of the force
distribution along the length of the weld need be applied, is 150 times the throat thickness, a.

Effective throat thickness, a

The effective throat thickness, a of a fillet weld may be taken as 0.7 times the leg length, s for a 90°
angle of fusion. The effective throat thickness, a, of a fillet weld should not be less than 3 mm.

The strength of a full penetration butt weld should be taken as the weaker of the connected parts,
assuming the weld material is stronger than the parent material. For a partial penetration butt weld,
the throat thickness should not be greater than the depth of penetration that can be consistently
achieved.

Design of welded connections to Eurocode 3

Design of butt welds

Details for the design of butt welds are as follows:

• Strength of butt weld taken as that of parent metal (i.e. fy in tension or compression or fy/ 3 in
shear) provided that suitable electrodes are used.
• Throat thickness taken as minimum depth of penetration, reduced by 3mm for most partial-
penetration butt welds.

Design of fillet welds

EN 1993-1-8 provides two methods for the design of fillet welds – the directional method and the
simplified method. The directional method considers all stress components acting on a critical
section of weld, and provides two conditions to be satisfied. The simplified procedure may be
applied independent of the direction of loading, and it must be demonstrated:

Fw,Ed ≤ Fw,Rd (9.1)

where : Fw,Ed is the design value of the weld force

Fw,Rd is the design resistance of the weld.

And the total design resistance of the weld may be calculated as follows:

Fw,Rd = fvw,d a l (9.2)


where fvw,d is the design shear strength of the weld (given by Equation 9.3)
a is the throat thickness of the weld
l is the length of the weld.

fu
fvw,d = (9.3)
β w γ M2 3

where: fu is the minimum ultimate tensile strength of the connected parts (see
Table 9.1)
βw is a correlation factor that depends on the material grade (Table 9.1)
γM2 may be taken as

Alternatively, evaluation of the weld may be made in terms of stresses, where the maximum
magnitude of stress on the weld σR must be shown to be less than the design weld strength fvw,d.

Table 9.1. Values for ultimate strength fu and correlation factor βw


Steel grade Thickness Ultimate Correlation factor
range (mm) strength fu βw
S235 t ≤ 40 360 0.80
40 < t ≤ 80 360
S275 t ≤ 40 430 0.85
40 < t ≤ 80 410
S355 t ≤ 40 510 0.90
40 < t ≤ 80 470
S450 t ≤ 40 550 1.0
40 < t ≤ 80 550

Example 9.1 – Welded joint design


Analysis of weld groups

Weld groups under eccentric shear

The bracket in Figure 9.7 is a similar configuration to the bolted bracket given in Figure 8.10, and is
subjected to a similar eccentric shear.

Figure 9.7: Eccentrically-loaded weld group

The analysis of this weld group proceeds by very much the same process as for the bolted bracket:
replacing the eccentric shear by a concentric shear plus torsional moment. The stress components
on the weld for these two effects are calculated separately, and then the maximum magnitude at
the critical location of the weld is determined.

As shown by the example in Figure 9.8, equivalent to the bolt group analysis, the following equations
may be used to determine the stress in the weld at the critical location:

M = R × c (where c = eccentricity of shear force from centroid of bolt group)

The shear stress τ carried by the weld due to the (concentric) vertical shear force R is:

R
τ = (where a = throat thickness of the weld; l = effective weld length)
la

Due to the moment, the following stress σm is induced at the critical location (i.e. furthest from the
centroid of the weld group).

M rmax
σm =
Io

Or, more conveniently in horizontal and vertical components:


M y max M x max
σh = and σv =
Io Io

where: Io = the polar second moment of area of the weld group = Ix + Iy

For example, the polar second moment of area for the shaded portion of weld of Figure 9.8, is given
by:

La3 a L3
I0 = Ix + Iy = + L a y12 + + L a x12
12 12

The distances x1 and y1 are measured from the centroid of the overall weld group to the centroid of
the individual portion of weld under consideration. The first term (La3/12) in the above formula is
commonly ignored because a << L.

Note that Io for the weld group (as required for the calculation of σh and σv) is obtained by the
summation of the values of Io for the individual portions.

Figure 9.8: Analysis of eccentrically-loaded weld group

The resultant stress σR is given by:

σR = (τ + σ v ) 2 + σ 2h

This must be demonstrated to be less than the design strength of the weld fvw,d.

Out-of-plane loads on weld groups

An example of a weld group subjected to out-of-plane loading is shown in Figure 9.9. The connection
design is simplified if the moment is small enough to be replaced at the joint by a couple in the
flanges without overstressing the flanges (i.e. F = M/d ≤ py×B×T). The welds attaching the flanges to
the column are then end fillet welds (or butt welds) of sufficient throat area to carry the force F. The
beam web welds are then simply carrying the shear force V as a side fillet weld. If M/d over-stresses
the flanges, then some moment has to be carried on the web welds by vector addition of the
bending stress with the shear stress. Note that the beam flange welds will have a limited effective
width if there is no stiffener at the back of the column flange.

Figure 9.9: Out-of-plane loads on weld groups

Example 9.2/Tutorial 9.1 – Beam to column welded joint design


10. INTRODUCTION TO BUILDING DESIGN

The layout and structural analysis of frames was introduced earlier in this course. The design of
buildings first requires an initial layout to be established, followed by a structural analysis, initial
member sizing, and lastly final member design and detailing. Although based on similar
philosophies, the design requirements and detailing of multi-storey buildings and industrial buildings
differ, and are therefore discussed separately herein.

Multi-storey buildings

Multi-storey steel frames fall into two principal categories – simple frames, where the joints are
‘nominally pinned’, and moment resisting, where the joints offer full structural continuity (refer back
to the different types of joint). In fact, an intermediate category also exists, where the joints are
semi-rigid. All frames must have adequate resistance to lateral loads; if a frame utilises simple
construction (i.e. pinned connections), it generally also contains a diagonal bracing system. Simple,
braced frames are a very common structural form for multi-storey buildings, and will be the focus of
this section.

Figure 10.1: Example of a multi-storey building


Beam design in simple frames

Beams in simple frames are designed as simply-supported and should be checked for bending
moment resistance, shear resistance and deflection requirements.

Column design in simple frames

Columns in simple frames may be assumed to be effectively pin-ended, so the effective (buckling)
length of the column may be taken as equal to the system length. Column bases in simple frames
are generally pinned. The joints between the beams and columns will be such that there is an
eccentricity from the beam reaction to the centreline of the column, and the eccentricity e is
typically taken as (H/2 + 100 mm), where H is the depth of the column. This eccentricity e gives rise
to a small bending moment in the column (M = R × e), where R is the reaction from the beam.
Strictly therefore, the columns in simple frames should be checked for combined compression plus
bending. For an initial design, it is acceptable to ignore the eccentricity moment.

Joints in simple frames

Schematic illustrations of typical types of joints in buildings are shown in Figure 10.2.

Figure 10.2(a): Beam-to-column fin plate joint


Figure 10.2(b): Beam-to-column end plate joint

Figure 10.2(c): Beam-to-beam end plate joint


Figure 10.2(d): Column base joint

Common practice in the UK is to assume that rotation will occur about the centreline of the
supporting beam or about the supporting column face, rather than at the connection to the web of
the adjoining beam. Thus connections at the centreline of the supporting beam or at the supporting
column face will have to resist vertical shear force only, whilst those in the beam web will have to
resist vertical shear, plus shear due to the eccentricity moment. If there is doubt about the position
of rotation, both cases should be checked. For end plate joints, such as Figure 10.2(b), eccentricities
are small and are generally ignored when determining bolt or weld forces.

Resistance to lateral loading (bracing)

The resistance of steel frames to lateral loading (generally wind loading) can be achieved by a
number of means, of which moment resisting joints, a stiff (concrete) core or diagonal bracing are
the most common examples. Combinations of these three systems may also be employed.

The use of rigid joints between the beams and columns offers total internal adaptability, with no
bracing between columns or walls to obstruct circulation. The disadvantages include increased
fabrication effort, larger bending moments transmitted to the columns, and generally a less stiff
arrangement than can be achieved with other bracing systems.

Reinforced concrete walls (shear walls) constructed to enclose lifts, stairs or services generally
possess sufficient strength and stiffness to resist lateral loading. Ideally, cores should be located to
avoid eccentricity between the line of action of the lateral load and the centre of stiffness of the
core arrangement, though this is not always possible.

Bracing systems in buildings act as vertical trusses which resist lateral loads by cantilever action.
Their positioning should be based on a similar principal to that for the concrete core, in order to
avoid the generation of additional moments due to eccentricity between the line of action of the
lateral load and the centre of stiffness of the bracing arrangement. Bracing members can be
arranged in a variety of forms, to act solely in tension or in both tension and compression. If
diagonal cross bracing is employed, depending on the lateral loading direction, one diagonal will act
in tension and the other in compression; by ignoring the contribution of the compressed diagonal,
the bracing can be designed to resist all the load in the tension diagonal. Typical bracing
arrangements are shown in Figure 10.3.

It is generally inefficient and impractical to brace all bays of a building, so bracing would typically be
employed in only one or two bays in each orthogonal direction, depending on the dimensions of the
building. Of course, there must be provision to ensure that the lateral loads can be transmitted to
the braced bays – this is often achieved through the members in the structural frame acting in
tension or compression or by the floors acting as diaphragms, but in some cases plan bracing will
also be required.

Figure 10.3: Bracing arrangements

In all cases it should be checked that there is sufficient lateral resistance in the structure carry the
total design lateral loading.
Bracing layouts
Calculation of bracing forces
Industrial buildings

There are generally two types of roof-supporting structures used in single-storey industrial buildings
which are portal frames and trusses. A truss system will give a lighter structure than the portal
frame. This is usually however at the cost of extra fabrication effort. Figure 10.4 shows an example
of an industrial building.

Figure 10.4: Example of a single-storey industrial building

Layout

A three-dimensional structure may be broken down into a series of two-dimensional components or


frames:

1. Roof sheeting (1.5-3 m) (typically continuous over 2 or 3 spans)


2. The purlins support the sheeting at 1.5 to 3 m centres. Typically now made of cold formed
sections spanning 6-8 m (6 m typical).
3. Plane trusses (or plane portals).
4. Columns supporting trusses.
Figure 10.5: Layout of building
Types of Roof Trusses

Figure 10.6: Common types of roof trusses: (a) Pratt – pitched, (b) Howe, (c) Fink, (d) mansard, (e)
Pratt – flat, (f) Warren, (g) modified Warren, (h) saw-tooth
Figure 10.7: Example of purlin and purlin cleat connection

Analysis of Trusses

Spacing’s of point loads on the truss are determined by roof sheeting span and may or may not be
applied to the truss nodes. Truss members are usually standardised to just a few section sizes for the
webs and chords. It is therefore not necessary to carry out a full truss analysis. Instead, the method
of sections may be used to identify the forces in the critical members.

For example, in the truss shown below which is standardised into three regions, only critical members
need to be considered. In the case shown below in Figure 10.8, verticals, diagonals and the two chord
members are checked in each region (though the two end regions are the same because of symmetry).
The analysis should ideally be carried out for unit loads, then member forces multiplied by W (load at
nodes) for each load case.

Figure 10.8: Analysis of a typical truss


Tension members in trusses

Apply normal design rules used for tension members. For single angles may need to consider
eccentricity. Double angles will usually be placed either side of a gusset plate hence no eccentricity is
applied.

Compression members in trusses

Overall buckling needs to be considered in both directions as a function of the truss layout and
connection details. Web members will not be continuous through a node and should in general be
designed for an effective length factor of unity. However, angles can be a special case. They may be
connected to a gusset plate (and hence have a reduced effective length). In single angles, attention
must be given to buckling about the minor v-v axis, whilst for double angles, iz is obtained from
parallel axis theorem).

Truss joints - general

Joints actually used in trusses do not reflect the idealisation of pinned joints assumed for simple
manual analysis. It would be expensive to make a truss with truly pinned joints. Experience has
shown that traditional joints give satisfactory performance.

The type of connection depends on the type of truss, the sections used for the members and
whether bolting or welding is to be used to make the joints. In general, truss joints may be classified
as follows:

Internal Joints:
These connect the truss members together and usually join discontinuous web members to a
continuous chord. Typical examples are shown in Figure 10.9. Members may be bolted or welded
(usually shop welding) and members may be connected directly or through a gusset plate. Where all
the members are angles then gusset plates will be needed at nodes. Chords will then usually be
doubles angles and webs will usually be single angles. Long-stem Tee sections (cut from UB) can be
used to weld web angles directly to the stem of the Tee (note that centroidal axes can be helped to
meet at a point by welding angles to both sides of the stem). As many as possible of the connections
will be carried out off-site in the fabrication shop where it is more economic to weld the joints.

It should be noted that in the case of gusset joints, making member axes meet at a point may lead to
a need for large gusset plates. These can be made smaller if the members are joined as shown in
Figure 10.10. In this case, the moment due to eccentricity should be taken into account.

Site Splices:
Large trusses must be sub-divided for transport to site where they are usually assembled through
site splices (usually bolted). Typical examples are shown in Figure 10.11. The gusset plate can be
used as the splice plate in light trusses. In heavily loaded trusses, cover plates are usually used to
connect the outstanding legs of the angles or the tables of Tee sections. In this case, the splice may
be located away from the truss joint. Site splices are generally bolted joints.

External Joints:
Cap or face joints are required to connect the truss to a column or another truss. Typical Examples
are shown in Figure 10.12. The members should be arranged such that the centre lines meet at a
point.

Figure 10.9: Internal truss joints


Figure 10.10: Joint eccentricity in trusses
Figure 10.11: Site splices
Figure 10.12: External joints

Design of truss joints

Bolted joints (see Figure 10.13):


Bolts in angles will usually be placed on the setting-out line, also referred to as gauge line (i.e. down
the middle of the connected leg). Although this will be at an eccentricity to the line of the force in
the angle, it is universal practise to ignore eccentricity moments in the bolts. Ordinary bolts are
usually used for truss connections. For design of ordinary bolts, divide force by number of bolts and
check shear (single or double) and bearing (bolt and plate).

Welded joints (see Figure 10.14):


In this case, the eccentricity between the applied force and the centroid of a weld group should not
be ignored. Instead of designing for an eccentrically-loaded weld group, the weld may be arranged
so that its centroid lies on the centroidal axis of the member (hence no moment effects and the
force may be divided uniformly over total weld length). Welds may be placed all round or placed in
the sides only. The first method seals against corrosion whereas the latter has some cost
advantages in terms of fabrication (but filler/sealant may be needed).

Gusset plates:
The thickness of gusset plate may be initially selected such that it is equal to or slightly larger than
that of the thickest part to be connected (usually minimum of 8 mm). The number of bolts or weld
lengths are determined for each member and the gusset plate is then made large enough to
accommodate the connections. Two approximate methods may be used to check the capacity of the
gusset. Simple elastic bending theory may be used to check critical cross-sections under axial load,
bending and shear. Direct stress may also be checked in the plate at the end of each member
assuming a dispersion angle of 30o.

Figure 10.13: Bolted connection design


Figure 10.14: Welded connection design
APPENDIX A: SECTION TABLES

Universal beams (UB)

Universal columns (UC)

Circular hollow sections (CHS)

Square hollow sections (SHS)

Rectangular hollow sections (RHS)

Elliptical hollow sections (EHS)

Channels

Equal angles

Unequal angles

You might also like