Vol 1. Structural Aspects
Vol 1. Structural Aspects
SWEDISH
WOOD
Design of timber structures
Structural aspects of
timber construction Volume 1
Volume 3
EDITION 3:2022
J SWEDISH
/If WOOD
: SWEDISH
8? WOOD
The book series Design of timber structures Volume 1-3 has been produced to make
it easier for structural designers to calculate timber structures and it is adapted
to Eurocodes. It is being used for higher education at universities and institutes.
The book series Design of timber structures Volume 1-3 includes Volume 2, Rules and
formulas according to Eurocode 5 as well as Volume 3, Examples. All three books are
available in English and Swedish. Since the books are available in both lan
guages and due to the nuanced content, our goal is that they will play a role for
many users on different skill levels. Notice that the Swedish decimal holder “,”
was kept throughout all the text, instead of the English “.”. All photos are taken
in Sweden, unless otherwise indicated.
Patrice Godonou
Swedish Wood
Contents
1.1 General introduction to structural design 7 4.2 Shear capacity of single dowels 89
1.2 Life phases of a building project 7 4.3 Johansen theory - timber-timber joints
and panel-timber joints 91
1.3 Various requirements related to timber structures 12
4.4 Steel-to-timber joints 95
1.4 Eurocodes - general assumptions,
limit states and detailed design 15 4.5 Expressions for the resistance of
a single dowel 97
1.5 Concepts used for the limit state design
of timber and wood-based products 20 4.6 Tensile capacity of single dowels - rope effect 100
2.4 Mechanical properties of wood and timber 33 4.13 Glued-in rods 111
3.5 Arches 86
Structural systems
for infrastructure 228
9.1 Timber bridges 228
Symbols 242
Bibliography 246
Introduction to design
and design process
Robert Kliger
Intuitive phase
This phase includes the definition of specific goals for the project,
understanding the client’s wishes and transforming them into a list
of requirements, understanding the design criteria as specified by
society via current design codes, generating simple concept solutions
for various ideas and the preliminary evaluation of each idea.
Understanding the client’s wishes and collecting all the necessary
information from possible stakeholders, which may affect the project,
is a crucial part of the intuitive phase. Literature studies, a search for
inconsistencies and interviews with users and other stakeholders
may provide the information necessary to define the aims and set
a list of priorities comprising different requirements. The design
requirements should be as clear as possible and should include issues
such as performance, quality parameters, reliability, safety, product
life span, aesthetics, ergonomics, economy and maintenance.
The main aim of the systematic search is to generate a wide range of
possible design concepts. A combination of intuitive and systematic
methods is probably the best way to generate new concepts. Creative
and innovative ideas are based on good theoretical and practical
knowledge. Various ideas should be represented by different struc
tural materials or combinations of materials, a variety of structural
systems and production methods. Such ideas are created by a system
atic search and are produced within the intuitive thinking process.
It is always advantageous to produce too many concept ideas rather
than too few. The main concern for a preliminary evaluation is to
reduce the number of concepts. This can be achieved by discussions
or by using preliminary evaluation matrices. It is important to have
a well-documented motivation of a solution that was made and per
haps be able to backtrack to this step in the following process. Strandparken, Sundbyberg.
Preliminary sizing
The aim of preliminary sizing is to estimate the preliminary dimen
sions of the load-bearing members in a building, such as trusses,
beams, columns, walls and floors. The thickness of the insulation
may be a decisive factor when it comes to the thickness of structural
members. The preliminary sizing often includes obtaining relevant
load combinations in order to estimate sectional forces acting on
a structural part which is being analysed.
Verification
Verification is the main part of structural design. The verification of
the load-bearing capacity of a structure or part of it is conducted in
the ultimate limit state, while the verification of whether it functions
is conducted in the serviceability limit state. To verify a structure in
the ultimate limit state, a designer needs to create structural and
load models by using design values for actions, material or product
properties and geometrical data. Load cases should be selected by
identifying load arrangements and possible deviations from assumed
directions and positions of actions and sets of deformations and
imperfections that should be considered simultaneously.
1.2.3 Construction
Process
The construction process usually involves the translation of designs
into the real production of a building. The design usually consists of
drawings and specifications, usually prepared by a design team
including surveyors, civil engineers, cost engineers (or quantity sur
veyors), engineers for electrician, structural design, plumbing and
fire protection.
The increasing complexity of construction projects is creating
a need for design professionals trained in every phase of the project
life cycle and the development of an appreciation of the building as
an advanced technological system requiring the close integration of
many subsystems and their individual components, including
sustainability.
Building Information Modelling (BIM) is the process of generating
and managing building data during the life cycle of a building.
Typically, it uses three-dimensional, real-time, dynamic building
modelling software to increase productivity in building design and
construction. The process produces a Building Information Model
Glulam frame assembly for the Infanterimuseum, Karlstad.
(BIM), which includes building geometry, spatial relationships, geo
graphical information as well as quantities andproperties of building
components.
A BIM is a way to handle information which is created during
the design and construction process but also during management,
that is all the way from conceptual design to construction and
through the service life. Modelling makes it possible to handle and
re-use the relevant information about the building.
1.2.5 Demolition
The demolition phase is the final stage in the service life of a struc
ture. Today, there is still a real opportunity for improvement and for
reducing the environmental impact of the building sector when it
comes to re-using and recycling building materials. The built
environment makes an important contribution to the total national
environmental impact. Government policies are expected to play
an important role in improving the environmental performance of
the building sector. The energy and material use in construction and
during the service life of a structure represent a large amount of
the total energy and material consumption. By using buildings func
tionally for as long as possible and by aiming during demolition for
the maximum re-use of materials and the adequate treatment of
the waste streams, a substantial environmental gain may be realised.
In this respect, timber structures may have a very large advantage in
comparison to other building materials. Demolition projects today
are more complicated if sustainability is incorporated into this stage.
The preparatory phase from policy to initiative must be well
described. Already during the conceptual design, decisions are made
which often have a major impact on the entire service life and sus
tainability of a structure, including the demolition stage.
Table 1.1 Matrix of requirements from stakeholders with examples at different levels, Johansson et al. (1990)
Building Comply with the design Function in specific terms. Aesthetic considerations, Choice of material
commissioner standards Dimension, shape, durability, “warm” surfaces, shape,
User smell, load-bearing capacity, easy to paint, repaint
stiffness, vibrations and dismantle
Contractor, builder, Stiffness Dimension, shape, stiffness, Right length, tolerant Species, discoloration
carpenter smell, load-bearing capacity, to moisture, no and appearance of
moisture content, durability, impregnation. built-in timber
mould, rot, wane, weight, Right price
location of knots, ability to
nail, screw and glue
Durability
The durability of a structure or a part of it assumes to have a certain
length of working life based on the fact that appropriate maintenance
is given and that it remains fit for use during the design working life.
The structural design is suitably adapted to the surroundings of
the structure or its parts. Load-bearing parts should be designed in
such way that deterioration will not impair the durability and perfor
mance of the structure exposed to the anticipated level of mainte
nance. Guidance for design with regard to working life of a structure
is given in Thelandersson et al. (2012).
Actions on structures
Geotechnical and
seismic design
d E.d
R.2> 1.2
This requirement has to be verified for each failure mode and a num
ber of different load combinations depending on the design situa
tion. The design value Ed of the load effect is determined on the basis
of permanent actions G, time-variable actions Q and accidental
actions A. The design value Gd of permanent actions is determined by:
Ga = rGGk 1.3
2d - Zq ' 2k 1.4
Td = 77— 1-5
/m
The factor y2 can also be regarded as a factor that converts loads with
short-term duration to an equivalent permanent action in order to
calculate the long-term deflection effected by creep.
EN 1990 defines load combination rules for the different design sit
uations, shown in table 1.2, page 16. These rules define how perma
nent loads and variable loads shall be combined to determine
the load effect. The following general format is, for example, valid
for design in persistent or transient design situations in the ultimate
state:
where indices j and i denote the jth permanent load component and
the ith variable load component respectively, and:
for accidental combinations and all combinations in the SLS, the par
tial factor yM is equal to 1. There are many different adjustments of
strength using modification factors. Commonly used factors in
Eurocode 5 are: kh,, kc,90,, kcrit,, kdis,, km,, ksys,, kv,, kvol . They correspond to:
Structural properties of
sawn timber and engineered
wood products
Marie Johansson
2.1 Forestry and the production 2.1 Forestry and the production
of sawn timber 23
of sawn timber 2.1.1
2.1.2
Europe 23
Swedish forestry 24
2.1.3 Production of sawn timber 25
In 2009, 3,95 billion hectares of land were classified as forest in
the world. Of these forests 25 percent were located in Europe, 21 per 2.2 Structure of timber 26
2.2.1 Material structure of wood 27
cent in South America, 16 percent in Africa, 15 percent in Asia,
2.2.2 Natural characteristics of wood 29
18 percent in North and Central America and 5 percent in Oceania.
Just under half of the wood use in the world is used as industrial 2.3 Physical properties of wood 31
wood (pulp, paper and wood products); the other half is used for fuel, 2.3.1 Wood and moisture 31
the main part of this in developing countries. The countries that pro 2.3.2 Shrinkage and swelling 31
2.3.3 Distortion of timber 32
duce the largest amounts of sawn timber are USA, Canada, Russia,
2.3.4 Density 33
Germany and Sweden. Canada, Russia and Sweden are also large
exporters of sawn timber. 2.4 Mechanical properties of
wood and timber 33
2.1.1 Europe 2.4.1
2.4.2
Strength and stiffness of wood 33
Strength and stiffness of structural timber 37
The forest sector is an important economic sector for Europe. 2.4.3 Influence of moisture 38
Studying Europe as a whole (including all of Russia) the forest sector 2.4.4 Influence of time 38
2.4.5 Influence of temperature 39
accounts for about 1 percent of the GDP and employs almost four
2.4.6 Influence of size 39
million people. Europe has 25 percent of the world’s forest resources
2.4.7 Long-term deformations 40
but only 17 percent of the total land area. The largest part, 81 per
cent, of these forest resources is within the Russian Federation. This 2.5 Strength grading 42
means that 44,3 percent of the European land area is covered with 2.5.1 Relationship between strength, stiffness
forests. In all countries in Europe the area and the amount of timber and other parameters 43
2.5.2 Machine strength grading principles 44
in the forest is increasing every year. The differences within Europe
is, however, very large. In most of the available statistics, Europe is 2.6 Modified wood 46
divided into three regions, Western Europe, Eastern Europe (incl. 2.6.1 Processes 46
the Baltic States) and CIS countries (former Soviet republics). 2.6.2 Properties of modified wood 46
The forests in Western Europe are mostly privately owned (over 2.7 Engineered wood products 47
70 percent) while in Eastern Europe the share of privately owned for 2.7.1 Engineered wood products
ests is increasing from having been mainly state owned before 1990. based on sawn timber boards 48
In the CIS countries the forest is to almost 100 percent publicly owned. 2.7.2 Engineered wood products
The fragmented owner structure in Western Europe has made it com based on veneers 50
plex to manage the forests and in many countries strong private for 2.7.3 Engineered wood products
based on strands, chips or fibres 51
est owner associations are formed. Forest management are in most
2.7.4 Built up structures - I-beams 52
European countries highly regulated by the government, meaning
strict regulations regarding harvesting and re-planting. The European 2.8 End-user requirement on sawn timber 53
forests (excl. Russia) are to a large extent a mix of natural and planted
forests. The forests in Russia are still to a large extent natural, espe
cially east of the Ural. Another important aspect of forest is the social
dimension. More than 90 percent of the European forests are open
for public access and the forest area for recreation is increasing.
The countries in Eastern Europe and the CIS countries before 1990
produced about half the sawn wood in Europe. Political changes
Figure 2.1 Example of two softwood species. a) Spruce (Picea abies), b) pine (Pinus Sylvestris).
deciduous species. The birch is mostly used in the pulp and paper Glued laminated
products Buildings
industry. 4 % cFurniture
new
Spruce and pine are used both in sawmills and in the pulp and 12 %
6%
paper industry. In 2013, 76 million m3 sub (solid volume under bark) Flooring
logs was brought to the Swedish industry, of this approximately 6% Renovation and
enlargement
50 percent was sent to the sawmilling industry and the other 50 per 8%
cent to the pulp and paper industry. The sawmilling industry yielded
about 70 percent of this into sawn timber and engineered wood prod Figure 2.2 Estimated use of sawn timber in Sweden,
ucts while the remaining 30 percent was wood chips sent to the pulp Dackling (2002)
and paper industry. The total volume of sawn timber from the saw
milling industry in 2013 was 17 million m3 of which 12 million m3
was exported. The largest export market for Swedish sawn timber in
2014 was Great Britain (2,7 million m3) with Egypt, Norway, Germany
and Denmark as next largest export markets (Skogsindustrierna 2014).
The use of Swedish sawn timber can be seen in figure 2.2, which
shows that a large part isused for packaging. Of the remaining sawn
material 55 percent is used directly in the construction industry.
S 3
60 90 °
S 2
5 15 °
S 50 70 °
P Random
21
Figure 2.5 The structure of a wood cell. The figure shows the cell wall Figure 2.6 Principle of the cell wall growth during one
constructed of 4 layers around the central cell cavity. Typical microfibril angle in year, note the thin-walled earlywood cells and the thick walled
these layers is given in the figure, Hoffmeyer (1995). latewood cells. The number of cells in a real tree is higher.
2.2.2.1 Knots
All trees need leaves (hardwoods) or needles (softwoods) for transpira
tion, respiration and the photosynthesis process. The area needed for
this process is so large that it is necessary for the tree to spread
the leaves and needles out onto branches. Each year a tree grows in
length at the top but also forms new branches at the sides. These
branches are each year gradually embedded in the stem wood
through the addition of more material on the outside of the trunk
and the branches; see figure 2.8. The material in the branch is
the same as in the main stem and the tissue systems of the two struc
tures are interconnected. A longitudinal section through a knot
shows that the fibre orientation around the knot is distorted and that
the fibres pass in a sweep around the knot. The presence of knots has
a negative influence on most mechanical properties of sawn timber.
The fibres around the knot are no longer continuous when the wood
is sawn and thus affecting the strength, at least in the main fibre
direction. The size, shape, number and location of knots influence
the strength of the sawn timber. For structural use, sawn timber with
smaller and fewer knots are often graded into higher classes.
Figure 2.8 A knot seen from all four sides of a wood specimen 45 x 45 mm in square section
m... - mArv
u = —----- *y ,100 2.1
Wdry
Figure 2.10 Relationship between equilibrium moisture content and the relative humidity
Figure 2.11 Diagrammatic sketch of micro fibril arrangement and shrinkage in normal wood and in compression
or juvenile wood. Note the difference in micro fibril angles.
a) Shrinkage at normal micro fibril angle.
b) Shrinkage at large micro fibril angle (compression wood or juvenile wood).
the microfibrils are in the S2 layer, the micro fibril angle of this layer
will govern much of the behaviour during shrinkage and swelling.
Since the microfibrils are normally inclined at a small angle to
the longitudinal axis, almost all the shrinkage occurs in the trans
verse directions. Juvenile wood and compression wood have microfi
bril angles that are much larger than those of normal wood, which
results in far larger longitudinal movements, see figure 2.11. Typical
values for shrinkage strain in the different directions are for a change
in moisture content of 1 % is:
• Longitudinally 0,0001
• Radially 0,0015
• Tangentially 0,0030.
2.3.4 Density
Density is a very important physical property of wood; it is correlated
to almost all mechanical properties. Density (p) is defined as:
m
P=V 2.2
where m is the mass (kg) and V the volume (m3). The density is mois
ture dependent as both the mass and the volume is dependent on
the moisture content. Therefore the density for wood has to be
defined also in terms of moisture content. The most commonly used
definition in timber engineering is the density p12, that is based on
the mass and volume at 12 % moisture content. This is the moisture
content used in all standard tests for wood strength.
The density of the cell wall is about 1 500 kg/m3. This means that
density is a measure of the amount of pores in the wood material.
The normal density (p12) for softwood grown in the Nordic countries
varies between 300 and 600 kg/m3.
Figure 2.13 shows the definition of the different stress in wood. For
timber the difference between R and T direction is often disregarded
and the directions are named g0 or g// and g90 or g± for the parallel and
perpendicular to the fibre direction respectively. To describe the “real”
behaviour of wood within the elastic range 12 constants are neces
sary; for example the modulus of elasticity EL, ER, ET, the shear mod
ules GLR, GLT, GRT and the Poisson ratios vLR, vRL, vLT, vTL, vRT, and vTR.
Normally, the Poisson ratios are assumed to be pair wise equal, which
make it possible to eliminate three. By disregarding the difference in
L - longitudinally tangential and radial direction, the number of variables can be
T - tangentially
reduced to six, often denoted E//, E±, G//, G±, v// and v where // denotes
R - radially
parallel to the fibre direction and ± perpendicular to the fibre direc
tion. The modulus of elasticity in wood is often shortened to MOE.
Figure 2.13 Definition of normal- (a) and shear- ( t)
stresses in different directions in wood 2.4.1.1 Tension parallel to the fibre direction
Loading small clear specimens (that is without knots) it is possible to
investigate the effect of loading the wood material in different direc
tions. A pure tension test in the fibre direction show that the stress
strain relationship is almost linear up to failure, see figure 2.14. On
the material level two failure modes are possible (or a combination of
them); the middle lamella breaks and the fibres are being pulled out
of their matrix material or the fibre can break. The strength of wood
in tension parallel to the fibre direction (ft) is very high; the failure
stress is often of the magnitude of 100 MPa. The failure is, however,
often very brittle.
zx o
a)
high some fibres will start to buckle and be driven into the other
fibres. When the buckling behaviour in the wood starts the possibil
ity to take higher load will diminish and the behaviour can be classi
fied as plasticising. Figure 2.16 shows the stress-strain relationship for
clear wood loaded parallel to the fibre direction. The compression
strength for wood in pure compression parallel to the grain (fc) is
around 80 MPa. Typical values for sc are 0,8 — 1,2 % with an ultimate
strain level su ~ 3sc.
Figure 2.18 Definition of the angle a between the stress (force) and the fibre
direction
TTL
Figure 2.22 The relationship between relative bending strength and loading time
The strength at 5 minutes has been set to 100 percent.
the same order of magnitude as for clear wood. There is also an effect
of moisture content and temperature on the DOL curve. Higher mois
ture content and varying moisture gives larger DOL-effect, the same
has been found for increasing temperatures.
To incorporate the effect of DOL in the design all loads are nor
mally classified into classes depending on their duration. Depending
on the load duration class of the specified load, the strength is
reduced. The reduction is larger for long term loads than for short
term loads. In Eurocode 5 the effect of duration of load and moisture
content is combined into one modification factor kmod, cf. section 1.5.1.
The volume is here the product of the width, depth and length, that
is b, h, and l of the specimens why the equation may be rewritten as
equation 2.6 with exponents associated with each dimension:
(4WkT
* ‘.iAr‘ .kt*' 2.6
• Knots — timber beams with more knots (and lower MOE) exhibit
larger creep deformation under load.
• Moisture content — tests have shown that wood with higher mois
ture content exhibit larger creep deformations than wood with
lower moisture content. Varying moisture content in combination
with loading have, however, been shown to give a significant
increase of creep deformations.
• Temperature — increased temperature leads to increased creep
deformations. This is especially noticeable at temperatures over
100 °C when the lignin is softening.
Figur 2.24 The deflection of small beech specimens loaded in bending to 1 / 8 and 3 / 8’s of
the short term strength in constant and varying climatic conditions, Hearmon and Paton (1964)
Measured parameter R2
Knots 0,16 - 0,27
Figure 2.26 The relationship between the modulus of elasticity (edgewise) and bending
strength for 380 studs of the dimension 45 x 70 mm
2912 29^2P 29
2.6.1 Processes
Acetylation is a reaction between acetic anhydride and hydroxyl
groups that are located mainly on the hemicellulose and lignin poly
mers located in the cell wall. The bonded acetyl bulks the cell wall
almost back to the green dimensions of the original wood. The reac
tion is carried out in a vacuum/pressure cylinder after the wood is
dried to below 10 % moisture.
Figure 2.28 Approximate year of development of different types of EWPs, beam type products above the time line
and panel type products below the time line. LVL - laminated veneer lumber, MDF - medium density fibreboard,
OSB - oriented strand board, PSL - parallel strand lumber, X-Lam - cross-laminated timber.
Strength
Figure 2.31 Distribution function for the strength of glulam beams and structural timber
Figure 2.34 Laminated-veneer lumber - LVL, all veneers with the same fibre orientation
2.7.2.2 Plywood
Plywood was one of the first EWPs to be produced. It is produced in
much the same manner as LVL, but with the veneers laid up perpen
dicular to each other. The number of veneers (plies) are always odd,
meaning that the fibre direction of the outer layers are always in
the same direction; see figure 2.35. Structural plywood is most com
monly produced in sheets of the size 1 200 x 2 400 mm or 1 220 x
2 440 mm with a thickness of 12 — 24 mm. The outer veneers nor
mally have the fibre direction oriented in the long direction of
the panel.
The structural properties of plywood are dependent on the number
and thickness of the veneers and load direction. It is important to
keep in mind the direction of the forces, bending perpendicular to
the plane of the panel and in-plane bending, see figure 2.36. The layers
with grain direction in the same direction as the normal stresses have
the highest stiffness and take the major part of the load. The contri
bution to the load bearing of the layers with grain direction perpen
dicular to the direction of the normal stresses is very small and can
in practice be disregarded.
a) b)
Figure 2.36 Bending a) perpendicular to the plane and b) in-plane bending including distribution of normal stresses.
The strands are in the process oriented to form the best possible
product, this means that in the outer layers the strands are oriented
parallel to the long direction of the panel and in the inner layer with
the strands randomly oriented.
The most commonly used OSB product is in the form of panels with
a size of 1 200 x 2 400 mm with a thickness between 6 and 25 mm
used as sheathing material in walls or in floor structures. OSB can
also be manufactured in larger panels with width up to 3 m, length
up to 25 m and thickness up to 75 mm. These panels can be used as
structural elements in the same manner as CLT panels.
Design of structural
timber elements
in ultimate limit state
Roberto Crocetti | Annika Martensson
b)
Figure 3.1 Examples of flexural members. a) External timber structure, b) floor joists.
The elastic theory of bending states that when, for example, a solid
rectangular member as shown in figure 3.2 is subjected to a bending
moment M about the y—y axis (the strong axis), the design stress at
any distance z from this axis will be:
M-z
3.1
WyRd ^Rd
3.3
k+
^yRd ^zRd
Values for the factor km are given by Eurocode 5: For structural timber,
LVL and glued-laminated timber, km = 0,7 for rectangular sections,
and km = 1,0 for other cross-sections. For other wood-based structural
products, km = 1,0 for any cross-section.
When a beam is subjected to bending, shear stresses will also arise.
In accordance with elastic bending theory, shear stresses will be gen
erated parallel to the longitudinal axis of the beam. The value of
the shear stress at any level in the cross-section of a beam, as derived
from elastic theory, is:
vs 3.4
I-b
where:
t is the shear stress at the required level.
V is the shear force at the position being considered.
S is the first moment of the area above the shear stress level
about the neutral axis.
I is the second moment of area of the cross-section
about the neutral axis.
b is the width of the cross-section at the shear stress level.
At any position along the beam the shear stress at the top and bot
tom faces of the cross section will be zero and the maximum shear
stress will arise at the neutral axis position. For a rectangular section
of width b and depth h the maximum shear stress will occur at mid
depth and will be:
3K
T =-------- 3.4 a
2b-h
This means that the shear capacity is given by:
V 3.5
V'Rd- 3
where:
A is the cross section area.
fv,d is the design shear strength.
3.1.2.1 Tension
Although the tensile strength, ft,0,k, of clear wood samples is greater
than the compression strength, fc,0,k, because tension failure occurs in
a brittle rather than in a ductile mode and also because of its sensi
tivity to the effects of grain slope, knots and other defects, the tensile
strength of structural timber is generally less than the compression
strength. This is particularly the case for the lower strength classes.
Design with respect to tension is normally very simple and the main
consideration is the strength value of the wooden member. In some
cases it can be relevant to make considerations of the volume sub
jected to the tensile stresses since it has been shown that the strength
is size dependent, see section 2.4.6, page 39.
In some cases the wooden members will be subjected to both bend
ing and axial loading. In the case of combined bending and axial ten
sion the design criteria are quite straightforward, that is failure will
not occur when the following relation is valid:
3.1.2.2 Compression
Examples of members subjected to compression forces are given in
figure 3.3 (columns, posts, stud members in walls or struts in trusses).
These are members that are subjected to a compressive action acting
parallel to the grain and along the centroidal x—x axis of the member.
When subjected to an axial load, as the slenderness of the member
increases there is a tendency for it to displace laterally and to eventu
ally fail by buckling. Since the main focus of members subjected to
compression parallel to grain is on buckling stability, compression
will be discussed in total in section 3.2, page 62. The same applies for
members subjected to compression and bending, it will also be dis
cussed in section 3.2.
when a stud is situated on a rail, so that only a small part of the rail
is subjected to compression forces. Consider for example the case
shown in figure 3.5 d), page 58. If the rail rests on the whole bottom
surface, the stress will be less at that boundary than at the top
boundary, since the loaded area is larger there. It should also be noted
that the stressed area on the top also can be considered to be larger
than the cross section area of the stud, since the fibres in the rail in
the vicinity of the stud increase the resistance against the force.
The latter factor can be taken into account either in the factor kc,90 or
by dividing the force with a larger area than the cross section area of
the stud. This would mean that ac,90,d„„. can be determined either by
N/bh or by N/Aef where N is the design bearing force, b is the width of
the stud and h is the depth of the stud and Aef is the effective contact
area, which then is larger than bh.
It is in many cases important to estimate the deformation in
the zones subjected to compression since in most cases it is the defor
mations and not the actual failure that will be of largest interest.
This is of larger importance in multi-storey buildings where
the deformations will add up from each floor.
a) b) c)
I
e)
Here, the problem with holes and notches will be discussed. Both
holes and notches may significantly reduce the load bearing capacity
of a beam and should preferably be avoided in design. Since tensile
forces are induced around a hole or a notch the failure will be brittle
and thereby sudden which means that there will be nearly no warning
before failure as there can be with other type of failures. In figure 3.7
the distribution of tensile stresses near the tip of a notch is presented
and it is obvious that the stresses can be larger than the tensile
strength. The theoretically calculated elastic stress will actually go
Figure 3.7 Stress at the tip of a notch according to linear theory and as
estimated in practice, full line shows theoretical stress and dotted line
estimated stress
towards infinity. Shear stresses may also lead to failures in the vicinity
of a sudden change in cross section area.
As the very high stress is often concentrated in a very small region
it is difficult to try to determine the load bearing capacity of a beam
with a notch or a hole with a conventional stress criterion. Instead it is
necessary to rely on tests or on concepts based on fracture mechanics
instead of conventional stress criteria.
Fracture mechanics is a part of the science of the strength of mate
rials. A solid body responds to extreme loading by undergoing large
deformation or fracture. Often the main interest for the engineer is
the magnitude of the load that causes fracture. In cases when there is
no or only minor stress concentrations, for example in the case of
a structural member in homogenous tension or bending, the calcula
tion of the fracture load can be carried out by a conventional stress
criterion. On the other hand in the case of a very high stress concen
tration some other approach is needed. Then different analyses can
be carried out applying theories of fracture mechanics. There are
a number of different alternative theories but none of them will be
described thoroughly here, only the basic ideas for the theory that
constitutes the equations given in Eurocode, namely linear fracture
mechanics especially analysis of the energy release rate. In this theory
the energy release in the fracture process region is the critical value,
that is when the energy release has reached a critical value Gc frac
ture will occur. The critical value can be seen as a material parameter
that can be determined by tests. By using Gc the value of the fracture
load can be determined, Gustafsson (1995). Other parameters that are
necessary to know are geometrical characteristics for the notched
beam as well as the modulus of shear stiffness and the modulus of
elasticity parallel to grain since those affect the behaviour. In
the equations of Eurocode it has been assumed that t/(BGc) is propor
tional to the shear strength, which has been shown to be the case
from tests.
In the design process the shear stress is controlled against the shear
strength in the following form:
where:
kv is a reduction factor that depends on the geometrical
conditions and material.
fv,d is the design value of the shear strength.
Vd is the design shear force.
b is the beam depth.
h is the beam height.
a is a non-dimensional parameter that is used to describe
the geometry of the notch.
i iv
■^ \ a I tz \ a
Figure 3.8 Notations for regulations concerning hole geometry and placement
In table 3.1 and figure 3.8 examples of these are presented. It is gener
ally recommended to place holes in the neutral axis of the beam.
Since there are large uncertainties in the few existing design rules,
the best thing at the moment is to assume that in areas with holes
practically no forces can be transferred by the timber but instead
reinforcement has to be used around the holes. This can for example
be done by reinforcing with rods or with plywood. In order to do so
there are some design advices given in literature that can be adopted
but also these should be used with great care.
lv > h > h
r > 25 mm > 15 mm
i
where the radius of gyration about an axis i = ,iyA), I is the second
moment of area, and A is the cross-sectional area of the member.
Buckling will occur about the axis with the highest slenderness
ratio. The effective length le (or buckling length) of a compression
member is the distance along its length between adjacent points of
contra-flexure, where the bending moment in the member is zero.
Examples of values for the buckling lengths that can be used in prac
tical design are given in figure 3.9. The values in figure 3.9 are slightly
higher than the theoretical values given by Euler theory because con
sideration has to be taken to uncertainties in boundary conditions.
For an idealized perfectly straight column of length l having uniform
properties and pin jointed at both ends, the theoretical axial load at
which buckling will occur about the y — y or the z—z axes will be
the Euler buckling load for the respective axis. Expressing the Euler Figure 3.9 Effective buckling length le for different end
buckling load in terms of the effective length, it can be written as: conditions. l is actual column length.
3.9
3.11
exceed unity. With the plastic theory solution however, the axial
stress to axial strength ratio becomes a squared function enabling
the member to take a higher value of bending stress to bending
strength ratio for the same value of axial stress to strength ratio.
In the case where no strength reduction is occurring, that is when
the slenderness is low in the member, the strength benefits associ
ated with the plastic behaviour of timber when subjected to compres
sion stresses can be adopted. This is described in the following equa
tion in the Eurocode 5 for the case with bending about both axes:
My^ | ^z,Ed [
3.13
,Rd Jid k ^c,O,Rd >
Horisontal
displacement
Vertical
displacement
Rotation
dw
3-15
dx
3.16
31 3 31 • I • 318 3.18
z z dx2
and:
3L.=G-K-30 3.19
dx
where:
E is the modulus of elasticity of the beam.
Iz is the moment of inertia about the weak axis.
Kv is the torsional stiffness factor.
G is the shear modulus.
Inserting equation 3.15 into equation 3.19 and then differentiating with
respect to x gives:
G-^v-^ = Mcrit-^
dx dx
3.20
321
or:
where:
J8(x = 0) = 0 => A = 0
M7 = -^E-Iz-G-Kn 3.27
ffmcr,t=
m,cnt — v 6T
Wy = b-h2-t ^48h
z GKvy 3.28
where:
m,dA
is the design bending stress.
fm,d is the design bending strength.
k crit is a factor which takes into account the reduced
bending strength due to lateral torsional buckling.
3.30
. _My,crit_ 6-7T
'0,05 ’ h ‘ ^0,05 ' Aor 3.31
m,crit W, bh2l
where the index 0,05 indicates 5th-percentile values of modulus of
elasticity and shear modulus (note that characteristic values are nor
mally adopted when stability checks are performed).
For a rectangular section of width b, depth h, the moment of iner
tia about the weak axis is Iz = (h • b3/ 12), and the torsional stiffness
factor for relatively slender beams can be taken to approximately
equal Itor ~ (b3 • h/ 3). Inserting these relationships in equation 3.31
the elastic critical bending stress can be written as:
7Cb2
'0,05 ’ ^0,05 3.32
= ITT
For softwood rectangular sections, the ratio of E0 05 / G0 05 is
approximately 16, and by applying this to equation 3.32, the critical
bending stress of a beam bent about its strong axis can be written as:
0,78-fe2 r
^”m,crit ~ h ^0,05 3.33
where:
E0,05 is the 5th-percentile value of the modulus of elasticity
parallel to the grain.
b is the width of the beam.
h is the depth of the beam.
l is the design span of the simply supported beam
between lateral supports at the ends of the beam.
0,78-fr2
C'm,crit 7 g ^0,05 3.34
Table 3.2 Effective length as a ratio of the span and loaded at the centroidal axis. If the load is applied at the com
pression edge of the beam, lef should be increased by 2 • h and should
Beam type Loading type lef /l
be decreased by 0,5 • h for a load at the tension edge of the beam,
Simply Constant moment 1,0 where h is the depth of the beam.
supported
Once the slenderness ratio for bending Arelm = fmk / gmcrit) 0,5 is known,
Uniformly distributed load 0,9 the reduction factor kcrit can be determined according to table 3.3.
Concentrated force at midspan 0,8 As can be seen from figure 3.16 and table 3.3, the value Arel,m = 0,75
Cantilever Uniformly distributed load 0,5 has been adopted as the limit below which the beam will be stiff
Concentrated force at 0,8
enough to prevent lateral torsional buckling. In other words, full
the free end bending strength can be achieved without risk for instability prob
lems if Arel,m < 0,75. This has been shown also experimentally, see
Source: Table according to EN 1995-1-1:2004, 6.3.3.
Piazza et al. (2005).
When Arelm > 1,4, that is when fmk > 2 • gmcrit the beam is considered
k
Table 3.3 Values of crit as a function of A rel
to fail solely by elastic buckling and the elastic critical bending
Value of kcrit. Relative slenderness strength will become the design condition. Finally, for relative slen
for bending, rel m A derness values between 0,75 and 1,4 the beam will fail in bending,
1 0,75
Arel,m < but after rather pronounced lateral (and vertical) deformations. Since
1 ,56 - 0,75 A rel,m 0,75 < A < 1,4
rel,m
there will be an interaction of bending in two directions (vertical and
lateral) and also torsion, the beam will not be able to achieve the full
) )2
1 / (*• A rel,m' A > 1,4
rel,m
design bending strength. Eurocode 5 adopts the approximation of
A graphical representation of kcrit plotted against the relative a linear relationship between these limits.
slenderness ratio for bending, A is shown in figure 3.16.
Figure 3.16 The value of kcrit for different values of the relative slenderness
ratio 1 rel,m
a)
c)
are transversal to the frame or to the arch, as shown in figure 3.17 c).
The second major problem with respect to the behaviour of frames
and arches in the lateral direction is that of lateral buckling (or out-
of-plane buckling). Since timber elements can be fairly slender,
out-of-plane buckling of the type illustrated in figure 3.17 b) may
occur.
One solution to the out-of-plane buckling is to increase the stiff
ness of the frame or the arch in the lateral direction by increased lat
eral dimension. When braced according to figure 3.17 c), the frame or
the arch can become fairly slender. It should be noted that the same
system used to stabilize arches or frames from overturning laterally
also provides lateral bracing and prevents lateral buckling from
occurring.
In normal portal frames and arches, out-of-plane buckling is checked
in the same way as any other beam-column between lateral restraints
and torsional restraints provided by bracings. These bracings make
the effective lengths of each element easily identifiable. In the case of
a portal frame, for example, the out-of-plane buckling should be
checked in the unbraced zones as for a column with buckling length
according to figure 3.18.
z'
h
Figure 3.18 Buckling lengths for out-of plane buckling of a portal frame
h/I
Figure 3.21 4 values as a function of
y h /1
72 Design of timber structures - Volume 1
3.2 Design of slender members
3.36
?crit’^2 EI
Wcrit =
8f
----- = V34------ 3.37
can be compared with the critical load for an axially loaded column
with length s = larch / 2 (= 0,53 • l):
»t El El
0111
2
AL-i, = 7T • —-----
(/Ls)2
3 = 32 ■ —j852---------
-0,532^2
3--3 3.38
3 = 1,17 3.39
^e = l,25-.S 3.40
= 1,25-a 3.41
h
Figure 3.23 Buckling length for in-plane buckling of a portal frame
xjz x^ x^ x^ xL xL x^ xL
a)
b)
0,0025^
Figure 3.25 Examples of assumed
initial deviations in the geometry for
a) frames and arches, when
b) symmetrical load and
c) non-symmetrical load. c)
Figure 3.28 Bending stresses (am), shear stresses (t) and perpendicular to the grain tensile stresses
(at90) in double tapered beam (left) and single tapered beam (right)
Figure 3.29 Stresses at the tapered edge of a beam. Left: bending stress parallel to the tapered edge
(principal stress). Right: bending stress a0 parallel to the grain.
Figure 3.30 Left: compression perpendicular to the grain occurs at the tapered edge.
Right: tension perpendicular to the grain occurs at the tapered edge. Nowadays the last mentioned
beam shape is not advised. It is from several aspects better and more common to use so called
fish beam, with curved laminations at the bottom surface and horizontal sawn at the top.
. 2
(T90 =CTma-Sin (Z 3.43
Equations 3.42 - 3.44 show that — close to the tapered edge — there
are both stresses perpendicular to the grain (o90) and shear stresses
parallel to the grain (t). The magnitude of such stresses increases with
increasing slope of the tapered edge. If the tapered edge is located at
the compression side of the beam, compression perpendicular to grain
occurs. On the other hand, if the tapered edge is located at the tension
side of the beam, tension perpendicular to grain occurs. For this reason,
tapered edges with an angle exceeding 10° on the compression side or
2 — 3° on the tension side should be avoided (it should be remembered
that tension perpendicular to grain can cause brittle failure at very
low level of stress and should therefore be kept as low as possible).
Since both the moment and the depth vary along the axis of
the beam, maximum bending stress as a rule occurs not where
the moment is greatest but at a section nearer the supports, see fig
ure 3.31. The position of this section can be determined analytically
from the condition:
rf(T(x) _ d
3.45
dx dx W(x),
x----- 3.46
from the support, where h0 is the depth of the beam at the support
and hap is the maximum beam depth.
Figure 3.31 Left: bending moment tending to increase the curvature of the glulam member. Right: bending
moment tending to straighten the glulam member.
distribution, it can easily been shown that the resulting tensile and
compressive forces, T and C respectively, give rise to a force T90 in
the radial direction:
d-&
3.48
Resultant forces T and C are equal to each other. For a beam cross sec
tion with breadth b and depth h, the resultants can be easily derived:
, h
C = -T = — Om'b'2 3.49
2
By geometric considerations:
d^ = rd'& 3.50
, h dt
<Tm b - 3.51
2 m 2 r
Knowing the tension force perpendicular to the grain, the corre
sponding tensile stress is:
T™ h i
a^=T^r~A-- =
bdl 4-r p
3.52
Equation 3.52 shows that the tension stress perpendicular to the grain
ot90 at the apex of a curved beam can be calculated approximately by
modifying the bending stress parallel to the grain (om = M/ W) with
a shape factor kp = h/ (4 • r). It is important to observe that increasing
the beam depth h and/or decreasing the radius of curvature r will
increase the magnitude of the stress perpendicular to the grain.
Various studies have shown that tension strength perpendicular to
the grain ft,90 is highly dependent on the stressed volume of the tim
ber. The basic design value of tensile strength perpendicular to
the grain must therefore be modified, for example by multiplying it
by a modification factor kvol and kdis:
3.53
where:
V0 reference volume. In EC 5, V0 = 0,01 m3.
V stressed/curved volume determined with regard to
the geometry of the member.
kdis modification factor with regard to stress distribution in
the beam.
Table 3.4 Values of kdis and V according to Eurocode 5 for typical beam types
Beam type kdis V
Double tapered beam 1,4 Volume of the stressed part (1)
(1)
0,5hap 0,5hap
a /(1)
ap
rin
r = rin + 0,5 hap
at x = x0 at x = t/2
Figure 3.34 Single tapered beam showing critical cross sections for bending stresses
At the outermost fibre of the tapered edge, the stresses should satisfy
the following expression:
where:
°m,a,d OCh °m,0,d are the design bending stresses at an angle to grain
and at the straight edge, respectively.
Md is the design bending moment in the section x = x0.
x = x0 is the position of maximum bending stress
(x 0 = (h0 • l)/ (2 • h ) for simply supported beams
with uniformly distributed load).
Wx0 is the section modulus at the cross section x = x0.
fm,d is the design bending strength.
k m,a is a reduction factor that takes into account
the simultaneous action of bending stress,
shear stress and compression/tension stress at
the tapered edge.
The values of kmma for different slopes of the tapered edge are shown
in figure 3.35. The values are derived for the glulam class GL30c.
The bending stress at the apex must also be checked, even though it
seldom governs the design:
. . ^aP»d _ l. ^^aP.d
^m,d *D ~ t\'6 n 3.56
W
'r ap bh 2
^^ap
where:
k is a factor determined by finite element analysis
that takes into account the tapering of
laminations. Values of k for glulam GL30c
are given in figure 3.37, page 83.
Map,d is the design moment at the apex.
Wap is the section modulus of the beam at the apex.
k
Figure 3.35 Values of ma according to Eurocode 5 for different slopes of the tapered edge, glulam class GL30c
■t,90,d -k
“ Kp w " Kp
. 6^ap,d 3.57
FKap ^ap
where:
kdis see table 3.4.
V see table 3.4.
ft,90,d is the tensile strength perpendicular to the grain.
m,a
hap
z b z
7T zr
Figure 3.36 Bending stresses and tension stresses perpendicular to the grain for: curved beam (left) and pitched cambered beam (right)
82 Design of timber structures - Volume 1
3.3 Special timber elements
Figure 3.37 Factor k according to Eurocode 5 for different radii of curvature, glulam class GL30c
_ j ^ap,d z 6Afap,d
<\90,d"V w * "
p ^2 3.61
ylap ‘'"ap
where kp is a factor determined by finite element analysis, defined as Table 3.5 Reduction factor for bending strength r k
the ratio between perpendicular to grain stress and bending stress at r /t,
according to Eurocode 5 as a function of the ratio .n
the apex. Values of kp for glulam GL30c are given in figure 3.38. r t
where in = inner radius of curvature and = thickness of
the lamination
The design tensile strength perpendicular to the grain shall be
reduced then in the same manner as for tapered beams, see equa rin /t kr
tion 3.58, page 82. > 240 1
k
p
Figure 3.38 Factor kp according to Eurocode 5 for different radii of curvature, glulam class GL30c
Design of timber structures - Volume 1 83
3.4 Portal frames
hr = h/15 + Z/30
hf ~ 0,7 • hr
r> 8 m
bh
3.62
joint" cos/?
w.. = bh2
3.63
Jomt 6-cos2/?
where:
b is the width of the frame cross section.
t is the width of the tips in the finger joints.
f is the spacing of fingers centre to centre at base.
3.5 Arches
Timber arches are generally made of glulam, mainly due to the fact
that glulam can be produced in curved forms and with varying depth
without a great increase in price. As a rule, solid sections of constant
depth are used, but composite sections of I or box form also occur (see
Chapter 5, page 115), especially for large spans.
The form of the arch should be chosen so that the moments are as
small as possible. As a rule, this means that the arch follows
the thrust line (equilibrium polygon) of the dominating loading com
bination. The influence of moments can however not be avoided
completely, as several load combinations must be taken into account,
each with its own thrust line. As a compromise a parabola is often
chosen. For functional reasons, for example to increase the headroom
near the supports, an elliptical or other arch form may sometimes be
preferable. The dividing line between frames and arches is fluid here.
The same result can be achieved by placing the arch on columns, see
figure 3.42, left. The horizontal support reactions caused by the arch
must in this case be taken care of by a tie rod between the springing
points of the arch. When the arch rests directly on the ground floor
slab, for example as in figure 3.43 right, the horizontal forces can be
taken up by the foundations if ground conditions permit, or by tie
rods under the floor or cast into it. To limit the size of the horizontal
reactions the rise of the arch should be equal or greater than 0,14 of
its span. For a parabola this corresponds to an angle at the base a of
30°.
For more parts, which are joined rigidly on the site, hinges and
rigid joints should be placed as in figure 3.43. The choice between
two- and three-hinged arches is made after similar considerations to
those for frames. Three-hinged arches are thus preferable over spans
of up to 60 — 70 m, while larger spans usually demand that the arch
is manufactured and transported in three parts.
f
Figure 3.42 Left: Arch with tie rod, on columns. Right: Arch springing from foundations.
Figure 3.43 Suitable placing of joints in arch structures. Left: Hinge, Right: Rigid joints.
the world. Dowel-type connections can be designed to be ductile, 4.8 Joints depending solely on
which ensures a safe structure. Dowel-type connections involve: tensile capacity 103
• Nails
• Screws 4.9 Brittle failure modes and group effects
• Dowels in dowelled joints 104
4.9.1 Group effect 104
• Nail plates (in combination with anchor nails/anchor screws) 4.9.2 Eurocode 5 application 104
and punched metal plate fasteners 4.9.3 Brittle failure modes in dowelled joints 105
• Bolts.
4.10 Forces acting at an angle to the grain 107
• Glued joints: can be used to connect structural elements, but also 4.11 Punched metal plate fasteners 107
to build up elements, for example glulam or LVL. Gluing requires
a controlled environment in production, since many glues are 4.12 Glued joints 109
4.12.1 Glue characterisation 109
affected by moisture content and temperature when setting. It is
4.12.2 Glue types 110
therefore not recommended to glue connections at a building site. 4.12.3 Strength 110
Glued connections often show a brittle behaviour, even though
the capacity might be high. 4.13 Glued-in rods 111
4.1 Dowel-types
a) Nails are smooth or rugged cylinders with a head, see figure 4.1.
The diameter is smaller than 8 mm. Nails can be driven directly into
wood or the hole can be pre-drilled. In Eurocode 5 the terms smooth
b)
nails and other than smooth nails are being used. A smooth nail is
a nail with a smooth shank without profiling, it can for instance be
Figure 4.1 Examples of nails round, square or grooved nails. To other than smooth nails, all other
a) Round, smooth nail nail types with a shank with some kind of profiling are counted, it can
b) Grooved, smooth nail (here a brad).
for instance be annular ring shanked, round jagged or twisted nails.
Screws are threaded cylinders with a head that usually do not
require pre-drilling, see figure 4.2. If the diameter of the screw is maxi
a)
mum 6 mm, it is theoretically treated as a nail.
Dowels are smooth (on rare occasions rugged) cylinders without
a head, see figure 4.3 respectively EN 14592 for dowel-type connectors.
According to Eurocode 5 dowels have a diameter 6 — 30 mm. All con
b) nectors to be used in timber structures must be CE marked by
the supplier.
Nail plates are combined with anchor nails or anchor screws and
c) have a pre-drilled pattern for the holes designed to automatically ful
fil the code requirements on spacing, see figure 4.4 a). Punched metal
Figure 4.2 Examples of screws plate fasteners (EN 14545) is a rationalisation of the nail plate with
a) Hexagon head wood screw requires pre-drilling nails, where ‘nails’ are punched and folded out from a steel plate, fig
and is however often replaced by for instance
ure 4.4 b). The plate thickness is maximum 2 mm to enable folding of
b) wood screw with countersunk head or
c) double threaded wood screw. the nails. They are pressed into the wood under controlled conditions,
for example when manufacturing timber trusses.
Bolts are screws that usually have a hexagon head and that require
pre-drilling, see figure 4.5. The diameter is for construction purposes
commonly at least 6 mm. Bolts often have a smooth shank and
a threaded part nearby the nut.
Figure 4.3 Example of dowel The material in dowel-type fasteners is most often steel, which can
be treated to become rustproof or hardened. Steel is characterised by
its ultimate strength, fu and yield strength, fy. A common system for
classifying bolts is to mark the fastener with a combination of num
bers for example 8.8 is translated to fu = 800 MPa and fy = 800 x 0,8
= 640 MPa.
Hereinafter a dowel referring to a dowel-type connector, such as
nail, screw, steel dowel and bolt.
40>k=0,082(1-0,01d)pk 4.2
/h,a,k = 4 . 43 * ----- 2—
*90 sin ct + cos a
[N/mm2]
L J
4.3
*90 = 1,35 + 0,015^
= 4.7
Jh,1,k
This situation can also occur if timber members of the same species
are mounted at an angle to each other implying that the embedding
strength may be different in the two members.
The first theory to establish the load-carrying capacity for a dow-
elled joint was put forward by Johansen (1949). He identified three
possible failure modes coupled to the number of plastic hinges form
ing in the dowel; failure mode I has no plastic hinges, failure mode II
has one plastic hinge forming in the dowel and failure mode III has
two plastic hinges, see figure 4.10, page 92.
t1 t2
For failure mode I the following relation is valid, see figure 4.11:
( 'I ( b2
4 I °2 ~
bright = /h2* I = Pfhi 02 ~ ”2" I
2 P
Expressing a1 = (t1 - b 1) / 2 and a2 = (12 - b2)/ 2
and substituting a second order equation remains:
F^ = fb^db1 = ^4k
V,KK ■'QJa -1 1+ 4.9
^=/h^(>/2-l) 4.10
4.11
v>Rk 2+p /h,l,k“^1
P /h,k^ 3MvRk
21 +----- ^-1
"v,Rk g
F 3M RL
_ fhX^2 2/1 +----- ^-1
V fhXdt2
When P = 1, it is seen that equations 4.11, page 93, and 4.12 describe
the same resistance.
P NWv+0
t1 t2
Using P = 1, the equation is reduced to:
Resistance single shear wood-wood connection with grooved nails, d = 4 mm andt1 =28 mm
1,5
0,5
0
0 10 20 30 40 50 60
Thickness t2 [mm]
Figure 4.15 Resistances in different failure modes for timber-to-timber single shear joint
Mode I
Mode II (d)
Mode III
Fvk,Rk
Figure 4.16 Resistances and failure modes for single shear steel-to-timber joint with fixed support of the nail from the steel plate
plate can be regarded as pinned and the dowel will merely rotate in
the hole without the formation of a plastic hinge. The resistance of
the joint is higher if a plastic hinge forms. Interpolation between
resistance formulas is permitted if needed. A full presentation of for
mulas for the resistance is found in tables 4.3 - 4.4, page 98 -99.
An example of failure modes and resistances is given in figure 4.16.
d
ff
^v,Rk - /h,2,k^2^
V1Rk
= 1.05
2+0 V
Al + 0) + + A- A
4/?(2AuA 4
/
d
Table 4.4 Single shear steel-to-timber joints, where tteel < 0,5d (thin steel plate)
5= =
DL
4
- 0,4/^ijtZ
l
4
<k
_________ p
----1
4
* ^=14572^4^+
1
4
'
Table 4.5 Slotted-in steel plates
d
r
^v,Rk =
■4-J ~ A----------------------- ______ L
+J-
p - f tA Io j. 4^y,Rk _ 1 . ^ax,Rk
* 'v.Rk Jh,l,kh« J2 + , ,.2 1 1 .
/hjjc"'! J 4 ■4-J
+J-
d
r ■
[
* ^v,Rk-2>37A?y-Rk^i-1’k^'1 4
/
4
i
4.6.2.1 Nails
For nails, the capacity in tension is dependent on the surface rough
ness along the nail and the anchorage capacity of the nail head. These
can be characterised through fax and fhead respectively, which can be
determined through tests according to EN 1382, EN 1383 and/or
EN 14358. If these strengths are not known, empirical expressions
exist to establish them for smooth nails (in other cases the nail is
tested according to EN 14592):
Expressions for calculating the withdrawal capacity Fax,Rk for the nail,
simply based on the surface area of the nail or the head, are then
given as:
■/axi^pen
^axjlk = min- (other than smooth nails) 4.16
■/headjc^h
^ixjc^pen
T^ax.Rk = min (smooth nails) 4.17
fax.k.dt "*■ -/headjc^h
where tpen is the penetration depth and dh the diameter of the nail
head. (For annular ring shanked nails tpen is the penetration depth for
the threaded part). Conditions apply to the thickness of the members,
see Eurocode 5, 8.3.2(7). Permanent axial loads are not allowed for
smooth nails. In figure 4.20, the addition of the withdrawal capacity
to the shear capacity is shown for grooved nails.
Fv,Rk
Fax,Rk
Fax,Rk /4
Figure 4.20 Withdrawal and shear capacity of grooved nails. Fv,Rk is the capacity in shear
and FaxRk the capacity in tension. The total capacity is the sum FvRk + FaxRk / 4. The joint is of
the same type as in figure 4.16, page 96.
4.6.2.2 Bolts
For bolts, the withdrawal capacity is dependent on the anchorage
capacity of washers and nuts and the tensile capacity of the bolt itself.
When determining the anchorage capacity of the washer, the charac
teristic compression strength perpendicular to grain for wood can be
taken as three times the characteristic base value due to concentrated
loading:
D = min- 12i
4J
steel
4.6.2.3 Screws
The withdrawal capacity of screws is dependent on the threading and
can reach very high values. The withdrawal strength can be determi
ned as:
4.19 4 0,52^’5^’1pk0'8 4 and /rf [mm], pk [kg/m3] 4.19
where a is the angle between the screw axis and the grain (a > 30°)
and kd is min(d/ 8;1) with d in mm, and n is the number of screws act
ing together in the connection. The outer thread diameter d must be
6 < d < 12 mm and 0,6 < d 1 /d < 0,75 where d1 is the inner thread
diameter.
The contribution from the withdrawal capacity for bolts (with
washer and nut) and screws is shown in figure 4.21.
Fvk,Rk
Fax,Rk (bolt)
Fax,Rk (bolt)/4
Fax,Rk (screw)
Fax,Rk (screw)/4
Figure 4.21 Shear and withdrawal capacity for bolts and screws. The absolute value of the withdrawal capacity is not
equal for bolts and for screws. The total capacity of the bolt respectively the screw in shear is taken as FvRk + FaxRk / 4.
2ax,Ed ,1--------
--------- 2v,Ed Si1 .. x
/(smooth.. nails) .4.21
^ax,Rd ^v,Rd
rax,Ed
(other types of nails and screws) 4.22
^^ax»Rd y <2v3d>
Tensile screw
Compressive screw
b)
0,10 -
0,05 -
Ihil
1st 2nd
Fastener position
nth
the highest load level, that is they will fail first, see figure 4.23.
The group effect is largest in joints with stout dowels, behaving
nearly linear elastic, see further section 4.12, page 109, on glued joints
where the uneven stress distribution causes the same type of problem.
The total capacity of such a joint is lower than the sum of the capaci
ties of the individual dowels. In joints with slender dowels, for exam
ple nails, the plasticity of the fasteners enables redistribution of loads
Figure 4.23 Load distribution in bolted joint,
so that the total joint capacity is close to or equal to the sum of
Lantos (1969). a) Force distribution in side member, the individual fasteners’ capacity, Blass (1990).
b) force distribution along the row of fasteners.
n
(screws d > 6 mm, bolts and dowels) 4.24
nef, (7d)
nef, (10d)
nef, (14d)
the capacity of the single fastener and the group of fasteners. Brittle
failures can be of different types; row shear, block shear, plug shear
and splitting, see figure 4.26.
If the joint consists of several rows of fasteners block/plug shear or
row shear failure may occur. For nailed joints, row shear never occurs
due to the small diameter of the fasteners. Furthermore, block shear
is uncommon since nails seldom protrude through the timber mem
ber. Nailed joints should therefore be checked only for plug shear
failure. Dowelled joints should be checked for block shear failure
according to Eurocode 5, but row shear failure is left out. Row shear
failure commonly shows a higher capacity than block shear, since
Figure 4.27 Faces along which splitting occurs in plug splitting occurs along more faces compared to block shear failure.
shear failure. In block shear failure no bottom face forms
The faces can be designated as in figure 4.27.
The area Anet,t is the net area at the end of the plug and the areas Anet,v
is the net area of the sides of the block. The factors 1,5 and 0,7 are
empirically derived factors to accommodate for two different phe
nomena. In the case of the tensile capacity, the tensile strength can
be higher locally at a joint, since the probability of having defects is
lower in a specified area than in a full size member, Larsen (2003).
The shear capacity is reduced due to a volume effect that affects
the shear strength when the loaded area is large, Johnsson (2004).
The areas should be reduced with the size of the fastener holes, as
indicated by the subscript “net” in equation 4.25. The method is
described in Eurocode 5, Annex A.
where Fv,Ed = max{Fv,Ed,1; Fv,Ed,2}, that is the largest of the two shear forces
in the beam on each side of the connection as shown in figure 4.28.
The design resistance is determined on the basis of the characteristic
resistance F90,Rk determined as:
Due to the production process, the punched metal plate fastener will
show an orthotropic behaviour. The strength of the fastener is depend
ent on the angle to the load direction and for a full characterisation
a) of a punched metal plate fastener no less than 11 strength parameters
Cantilever (recessed support) need to be determined; shear, tensile, compressive and anchorage
strengths in perpendicular directions and three constants which
describe the anchorage behaviour, EN 1075 and EN 14545.
When designing a joint with punched metal plate fasteners,
the tensile, compression and shear strength must be checked along
with the risk for buckling of the plate at unsupported sections of
b)
the assembly. The procedure is described in detail in Eurocode 5, sec
tion 8.8, but is seldom performed by hand calculations. Well developed
Punched metal software exists for the concurrent design of punched metal plate fas
plate fastener teners and trusses, taking into account the effect of movements in
the joints, which affects the force distribution in the truss.
Punched metal plate fasteners can also be used to reinforce timber
c)
itself, but also the area around joints, Nielsen (2003). In that case, it
is not the punched metal plate fastener in itself that is interesting to
design, but rather the effect it creates by preventing failure perpen
dicular to the grain in timber. Examples of usage are, see figure 4.31:
• increasing the compression strength perpendicular to grain by pro
viding a secondary load path in the connection and stiffening
the wood
• increasing the stiffness in bending by reinforcing the compressive
side of the timber member
• increasing the tensile capacity perpendicular to grain in the joint
area to prevent brittle failures such as block shear or splitting and
also provide high embedding strength at the most heavily stressed
section in a dowelled joint.
Reinforcement Reinforcement
c)
Figure 4.32 Classification of different types of structural adhesive joints, Aicher (2003)
Table 4.9 Characteristics of adhesives and comparison with wood, Aicher (2003)
fv GT
f fv2 / GfT f Gf. f2 / Gf.
Glue [N/mm2] [N/mm] [N/mm3] [N/mm2] [N/mm] [N/mm3]
4.12.3 Strength
The strength of a lap joint was first described by Volkersen (1938)
who arrived at the conclusion that the distribution of shear stress t(x)
in the joint can be described as:
Refer to figure 4.33 for notations. The coordinate x increases from left
to right along the length l.
b)
The most important result from the Volkersen theory is that the stress
distribution in the joint is non-uniform. The strength of the joint
cannot be determined by calculating tm, which is the average shear
stress value. There will be peak stresses at the ends of the glue line
where failure will be initiated. The same phenomenon, but less
prominent, is behind the group effect in dowels placed in a row.
A deeper study of the shear stress t(x) will reveal that above a certain
length, there is no change in peak stress or in other words, lengthen
ing the connection will not increase its strength. Volkersen theory
can be generalised to non-linear fracture mechanics, Gustafsson
(1987). By studying the limit cases for non-linearity, fully plastic bond
line and fully brittle bond line, Gustafsson (1987) arrived at upper
and lower bounds for establishing the capacity of a joint:
4.29
Pmm ~ ^/^(l +
If a fully brittle bond line can be assumed, the capacity is not influ
enced by the length of the bond line. However, the range for the brit
tleness ratio where non-linear fracture mechanics is valid is depend
ent on the bond length through:
0,1(1 + a)<co< 10 (1 + a)
^2/v2 4.30
(D =—:i2—
Z1E1GfT
For smaller a, plastic conditions apply, while for larger a, linear frac
ture mechanics should be used. Using l = 400 mm, t = 45 mm, and
E1 = 13 000 MPa and fv and GfT for a resorcinolphenol glue according
to table 4.9, page 109, a = 27. For a = 1 (symmetrical lap joint), the limit
value for a = 20, which means that linear fracture mechanics can be
applied. Changing glue to polyurethane makes a = 3,8, which means
that non-linear fracture mechanics must be used.
a)
Figure 4.34 Failure modes for glued-in rods, Tlustochowicz et al. (2010)
5. The term Sri2 is a characteristic of the joint and is called the polar
moment of inertia: Ip = Sri2 = S(xi2 + yi2)
6. The load component Fmi can now be established through:
F^-Kar,-^
2p
Figure 4.35 Load distribution in a moment resisting joint Myx Mx{
—- t
Fmxi■ =----tF ’ • = myi
•'p yp
^i=XariFy
i=l 4.31
We = aeF+
Vertical equilibrium: F - 4 • Fy = 0
Moment equilibrium: F • e - Fy • r = 0
y where Fy is the yield load for the fasteners in this case assumed equal
Figure 4.36 Example of lower bound establishment for all of them. The lowest value of F found from these two equilib
of plastic load distribution in joint, Larsen and Riberholt rium relations constitutes the lower bound solution found from
(1999) the assumed force distribution, that is:
F = min- 4.33
where pm is the mean density in kg/m3 of wood and d the outer diam
eter in mm of the fastener. In joints between wood and steel and
between wood and concrete, Kser established using the wood density
can be multiplied by 2.
By composite timber elements here is meant I-beams, hollow section 5.1 Structural elements
beams and stressed skin panels (SSP). As examples, the first two are with full composite action 116
typical beams having cross-sections as shown in figure 5.2, page 116, 5.1.1 Glued thin webbed beams 116
while the last one is a kind of building block covering a large surface 5.1.2 Glued thin flanged elements (SSP) 133
as in figure 5.14, page 133.
5.2 Structural elements
Composite timber elements are usually, but not necessarily, made
with partial composite action 142
of more than one material. Common wood based materials used are 5.2.1 Built-up bending elements 143
structural timber, glued laminated timber, laminated veneer lumber 5.2.2 Summary of equations 156
(LVL), plywood, fibreboard, particleboard, oriented strand board (OSB), 5.2.3 Failure criteria for built-up beams 157
etc. Other materials such as steel, aluminium, concrete, plastics, and
even glass are possible to use. But here we restrict ourselves to wood
based materials.
Basically these elements carry the load through bending action,
especially in horizontal structures like floors, roofs, bridge decks, etc.
Of course axial forces must be considered for structures affected by
such forces, like in wall elements and columns.
The common property of all these composite elements is that
the section is divided into wide outer flange parts with an intermedi
ate thin web part. Naturally, the flanges of an SSP-element are very
wide and very thin, but the web panels are still quite thin in relation
to the flanges. The same is also true for an ordinary I-beam, even
though the flanges may have a small width the web panel will have
a much smaller thickness. When these kinds of elements are subjected
to bending, most of the bending moment is carried by the flanges
while a minor part is carried by the web. The shear force on the other
hand is mainly carried by the web panel or panels. The only exception
is T-shaped sections for which no real distinction between a flange
and web part can be made.
Consider now a small cut-out from the compression flange of
an ordinary I-beam as shown in figure 5.1 a). The axial stresses change
slightly when moving from one end to the other end of the cut-out.
The only thing that can cause this change is the shear stress affecting
the web panel. All the shear stress must be transferred into the flange
in order to build up the axial stresses or vice versa. Now if the web-
to-flange connection is sensitive to shear deformation a small slip
will result between the two materials, as shown in figure 5.1 b). If this
slip is negligible we have so called “full composite action” (fca) in
a a + da Slip
a)
which case the beam’s stiffness and stress magnitudes can be based
on ordinary beam bending theory as explained in section 5.1. If
the slip is non-negligible the applied load will be counteracted differ
ently such that with increasing slip the flanges will carry less of
the applied bending moment through axial compression and tension
of the flanges and more through individual bending of each flange and
web panel. This will result in a decreasing transfer of shear stresses
over the interface. In addition to this redistribution from axial flange
forces to flange bending the stiffness will decrease leading to increased
deflection. This state of non-negligible slip is referred to as “partial
composite action” (pca) and is explained in section 5.2, page 142.
Typically, full composite action is usually assumed for glued con
nections having very small glue line thicknesses. Partial composite
action must be assumed if the glue is replaced by mechanical fasten
ers and/or the interface have a thickness such that its shear deforma
tion is not very small. Actually, the shear deformation of the web
panel itself can be treated as a slip and can be accounted for through
the theory explained in section 5.2, page 142.
Figure 5.3 I-beam or hollow section beam having: a) small web slenderness or b) large web slenderness.
Aspects 10 and 11, out of this extensive list of failure modes and
requirements to be checked, are of little or no importance for beams
of practical dimensions, and not treated in this book. The remaining
aspects must at least be thought of whenever you attempt to design
a beam of this kind. Usually some of them can be ruled out for a par
ticular problem. Detailed design rules and recommendations are given
in the following sections regarding most of the identified failure
mechanisms and serviceability requirements.
a)
cross-section is needed. The simple reason for this is that we must try
to predict the most likely distribution of stresses and avoid stress dis
tributions affected by large scatters in material data. Note, however,
that this reasoning applies to finding the best stress distribution
affecting a composite cross-section and may not be entirely correct
for instability problems, which is discussed in connection to equa
tion 5.4 below.
Eurocode 5 uses the concepts of instantaneous and final conditions
(“inst” and “fin”) to handle the creep problem. In a more general case
we have n parts where i = 1, 2...... n; and if one of the i-values is asso
ciated with the reference material, r, the fictitious width of the ith
part is:
r,mean
„ ^.mean (1 + ^2 V)
where "i,ULS,fm “ F T. ~ V 5.2
r,mean + V^2 ^i,def /
E. (1 + ^.Hrf)
i,mean \ r.def /
iUi,SLS,fin “ T T. 7 V
r,mcan + ^j;def /
Here ULS and SLS refer to the ultimate limit state and the serviceabil
ity limit state, respectively. ki,def is the creep factor of the ith member
as if subjected to permanent loading and specified in Eurocode 5 for
different materials. y2 is the reduction factor for quasipermanent
loading and should be determined for the load causing the largest
stress. If that is a permanent action, a value of 1 should be used.
When and how to use the different a.-values in equation 5.2 is
explained as:
«.._„.
i,ULS,inst
= V i,SLS,inst» is used in the ULS or SLS under instantaneous condi-
tions, that is as if all design loads are applied during a period
shorter than a year or so. This value should, in the ULS, be
used to check the resistance of those members that are most
prone to creep.
For global stability problems such as column buckling and lateral tor
sional buckling of composite beams, the recommendation to always
use the mean value of the elastic and shear moduli will be in conflict
to the definition of slenderness parameters (see sections 3.2.1, page 62,
and 3.2.2, page 64), which are generally defined as:
A:,rel “
Axial resistance
Critical axial load
^crit,rel ------- 5----------------
Bending resistance
Critical bending moment
5.4
The conflict is that the slenderness parameter should use the most
probable ratio, which is achieved if both the numerator and denomi
nator are based on either characteristic or mean values of
the involved material parameters. Only characteristic material prop
erties are available for the numerator (the resistance is based on fm,k,
ft,k, fc,k, etc.), while both characteristic and mean values are available
for the denominator (the critical load depends on E, G, etc. and not
on any strength parameters). It is therefore recommended that E- and
G-values used in stiffness expressions are taken as (lower) characteris
tic values, while the fictitious section is still based on mean values of
the moduli. Typical stiffness values involved in finding critical loads
are Er Iy,fic, Er Iz,fic, Er Iw,fic and Gr It,fic for y-axis bending, z-axis bending,
warping and torsion, respectively.
The difference is usually smaller than 5 percent in the final design
resistance of a typical composite cross section, if comparing results
obtained using either characteristic or mean values of the elastic
moduli involved in determining the dimensions of the fictitious
cross-section. If global instability is a part of the analysis the differ
ence usually stays within 10 percent, if mixing characteristic and
mean values when determining the slenderness parameter according
to equation 5.4. Hence, we conclude that a principal mistake regarding
the used moduli will have a small impact on the final result, if com
pared to the large scatter of material data and the magnitude of par
tial coefficients used. The same also holds if mixing up the i-values p
given in equation 5.2, page 119, which in general will have a small
influence on the final result. One can really ask if all the extra work
p
associated with these different i-values are worth the effort in practi
cal design situations.
^ aw,c,max,d
„
aw,t,max,d
Figure 5.5 Stress distributions on an I-section (or box-section) used to determine its bending moment resistance My,Rd
Note that the “real” stress distributions indicated is valid if Ew < Ef,c < Ef,t.
= £dot My,Ed | |
udot,d £ t pdotl 5.5
r y,fic
where zdot is the distance from the neutral axis to the dot and Er is
the modulus of elasticity of the reference material (in this case
the timber of the compression flange). It is now tempting to use equa
tion 3.6 in section 3.1.2, page 57, to verify the resistance of the flanges.
This is because the total stress affecting a timber flange can be split
into two parts: one pure compressive or tensile stress at the centre of
each flange and one pure bending stress, taken as the difference
between the outermost edge stress and the centre stress. Finally, sub
stituting these stresses into the appropriate interaction formula will
verify the resistance. It has however turned out that such an approach
is on the safe side compared to experimental results and also a bit
cumbersome. The recommended approach is much simpler. Usually
the flange depth is much smaller than the beam depth, in which case
their resistance is almost entirely controlled by the compression or
tensile strength of the flange material. Hence a reasonable design cri
terion would be to avoid stresses at the flange centres that are greater
than the material strength at these points. But this approach is too
optimistic if the flange depth becomes large in relation to the beam
depth. This obstacle is overcome by realising that the bending
strength is greater than the compression strength and much greater
than the tensile strength for most timber materials. So the problem is
solved by adding the additional requirement that the stress at
the outermost fibres of the flanges must be smaller than the bending
strength. For the dots in figure 5.5, page 121, the following six verifica
tions must be made:
where all g-stresses are calculated using equation 5.5; fc0d and fmd are
the compressive and bending strength of the timber in the compression
flange, respectively; ft,0,d and fm,d are the tensile and bending strength
of the timber in the tension flange, respectively; fc,w,d and ft,w,d are
the compressive and tensile strength of the web, respectively.
Using fc,w,d and ft,w,d will slightly underestimate the resistance of
the web panel/panels. The three main reasons for this are:
1) The strength values have been determined based on testing of
fairly wide and uniformly stressed specimens. But in a beam just
a small volume close to the edge is stressed to this level. 2) A lamina
tion effect in the flange, where the timber will allow for some stress
redistribution from the web into the flange before final web failure.
3) Web material of wood based sheeting has a tendency to break away
the surface material if subjected to compression. This is more or less
prevented for a surface that is glued to the timber. 4) The shear defor
mation of the web panel actually leads to a small increase of the axial
stresses in the web, which slightly counteracts the favourable effects
of remarks 1 to 3. By considering all these effects it should still be
possible to increase these particular strength values by 10 percent,
without hesitation. Such an increase is presently not promoted by
the Eurocode.
Note that the verification criterion (equation 5.6) does not cover
the risk of lateral torsional buckling. The background to lateral tor
sional buckling of ordinary rectangular timber beams are treated in
section 3.2.3, page 65. Here some additional comments regarding thin
webbed composite beams will be given. One very simple and easy to
use method is to consider the compression flange to be an ordinary
column that can buckle sideways without having to drag the rest of
the cross-section with it. The column is cut out between two points at
which the compression flange is sufficiently braced laterally. Pinned
boundary conditions are assumed at these two points, that is
the flange is treated as an Euler 2 column. Verification of the resist
ance is given by:
<7,..,
t,c,d 5.kf n.
c J c,0,d 5.7
My,cnt ent
~^y,Rk
A< 5.9
Vy My.cntv dim)'
Critical load
kcrit
kc glulam
kc structural timber
Figure 5.6 Buckling curves for lateral torsional buckling and flexural column buckling
The whole idea behind the chart offigure 5.6 is that the influence of
geometry, support and loading arrangements enters via A, while
the buckling curves only corrects the resistance with respect to all
relevant imperfections, where the most important ones for wood
based materials are non-linear material and bow imperfections.
The critical load is determined under the assumption of a linear elas
tic material that can sustain infinitely large stress. In reality kcrit or kc
can never exceed unity. We see that the two kc-curves give a slightly
greater reduction than the kcrit-curve. It can also be shown that
a curve for lateral torsional buckling must give smaller reductions
than for flexural column buckling. Since no specific curve exist for
lateral torsional buckling of composite I- and box-sections it is here
proposed that the present kcrit-curve can be used also for composite
beams, especially since most composite beams have smaller bow
imperfections than timber beams with rectangular cross-sections.
And if one for some reason feel uncertain about the applicability of
the kcrit-curve it should be safe to use the upper kc-curve for flexural
buckling of glulam.
The procedure to find the resistance with regard to lateral torsional
buckling is summarized as follows: The characteristic short term
properties should be used when calculating the slenderness accord
ing to equation 5.9, page 123, since the slenderness only reflects geom
etry, boundary conditions and load arrangement. But after determin
ing the reduction factor kcrit as a function of A the reduction should be
applied in equation 5.8, page 122, where My,Rd should be calculated
based on design values modified for duration of load with k mod
according to section 2.4.4, also see section 3.2 in Volume 2. It is recom
mended to determine the slenderness parameter A under instantane
ous (short term) conditions only, since a determination under final
conditions (long term) would have a small and insignificant impact
on the ultimate resistance. But note that in finding My,Rd itself both
“inst” and “fin” conditions may have to be considered.
bending failure caused by shear buckling. For web panels with open
joints the reader is referred to Thielgard and Larsen (1978).
Approximately, we may assume panel shear failure if hw /b w < 35 and
shear buckling failure if 35 < hw/bw < 70. Other kinds of buckling
phenomena may limit the resistance for values above 70. But web
panels having a slenderness above this limit are rarely used in practi
cal applications. Actually, a slenderness hw /bw above 70 should be
avoided unless the designer is experienced enough to handle all
instability scenarios that may occur. Note that the limit 35 is sup
posed to be valid for all board materials, but values adjusted to better
match the properties of individual board materials are given in
table 5.2, page 139.
Let us start with a pure panel shear failure under linear elastic con
ditions. The shear stresses counteracting the shear force Vz,Ed are then
distributed over the cross-section as shown in figure 5.7.
The simple failure criterion is that tvEd < fvmind. The shear stress can
be obtained using the well-known shear formula as given by:
^z,Ed ^y,fic(Z) V V
_ z,Ed z,Ed
Tv,Ed ~ £ 5.10
Iy,fic
. bw
where the left expression gives the shear stress parallel to the x- or
z-axis at any point z in the web panel. Sy,fic (z) is the static moment of
all area on either side of the z-coordinate, which is measured from
the centre of the fictitious cross-section. The only problem with this
formula, aside from its complexity, is that it overestimates the magni
tude of the shear stress by some 5 to 10 percent for ordinary I-beams.
The reason is that it is derived under the assumption that no defor
mations are caused by the shear stresses and that only axial strains
caused by the bending moment contribute. In reality shear stresses
will cause shear straining, which in turn will reduce the shear stress
below the value given by the left part in equation 5.10. If we instead
assume all shear stresses to be uniformly distributed over
the cross-sectional area of the web panel, Aw measured from flange
centre to flange centre, a much simpler relation is obtained. We have
in this simpler failure criterion neglected the slightly parabolic shape
of the stress distribution in the web panel. But this is justified by
neglecting the shear contribution from both outer flange parts. And
it is also justified by the fact that the actual shear stresses are smaller,
as described above. The simplified version of equation 5.10 is recom
mended for practical design, as in Eurocode 5. Note thatfvmind is
the minimum value of the panel shear strength taken parallel or per
pendicular to the beam axis. The shear strength of plywood varies
Tv(z)
Tv,Ed
y
considerably with the angle between its fibre orientation and the lon
gitudinal beam axis. The shear strength of plywood is more than twice
as large as the tabulated value for an angle of 45°. In Eurocode 5
fv,min,d is denoted fv,0,d and no distinction is made between panel shear
stresses acting parallel and perpendicular to the beam axis.
Figure 5.8 shows two side views and one end view of typical I-beams
for which shear buckling is limiting the resistance. The buckling pat
tern is clearly indicated. Note that the web panel must under all cir
cumstances be furnished with transverse web stiffeners at the sup
ports, whenever shear buckling is limiting the resistance. But more
stiffeners are rarely needed, unless large concentrated forces are
applied to the beam in which case additional stiffeners may be needed
at those locations. Note that web stiffeners may be needed at the sup
ports even if shear buckling is no problem. This is because of crushing
or local buckling of the web panel due to concentrated loading, that
is patch loading may limit the resistance. The best practice to avoid
patch loading failure where concentrated force are applied, like at
supports, is to always use web stiffeners at these locations. How to
check the patch load resistance if no stiffeners are used is not explic
itly covered in this book.
By rewriting the right part of equation 5.10, page 125, and introduc
ing a reduction factor xv covering the shear buckling phenomenon
we have:
bf
Flange
Stiffener
Web
Cross-section
h /bw 2 70
Figure 5.8 I-beams for which shear buckling is limiting the resistance, that is 35 < w
k
Figure 5.9 Buckling coefficient T for a simply supported plate subjected to uniform shear stress
around all four edges. a) Isotropic plate, b) orthotropic plate.
For isotropic plates, see figure 5.9 a), the buckling coefficient and its
corresponding shear buckling stress can be determined as:
where v is the Poisson’s ratio, which for hardboard can be set to 0,20
and to zero for particleboard if no better values are available. Note
that wood based board materials can be viewed as isotropic if
the fibre orientations are random within the plane of the plate (like
for particleboard and hardboard, but not for plywood and OSB which
have a clear grain direction). The buckling coefficient is simply a par
abolic approximation of the theoretically “true” result shown in fig
ure 5.9 a). For orthotropic plates it is a bit more complicated but
the buckling coefficient can be taken from figure 5.9 b), where param
eters fl1 and fl2 together with the shear buckling stress are given by:
5.14
Critical load
Design curve
Figure 5.10 Design shear buckling curve 0,8/2v , compared to the critical curve 1/2v2
Zv = min 5.16
7C2E
*3
1,25 'i
^-0,8
h
ZV=^O,8
•Ar,minjc v,mm,k
5.16
b
0,8 = 29,1—
hw
Point of
contraflexure
Vf
Vf/2
b) w /2
Figure 5.11 Local flange bending between transverse web stiffeners contributes to the shear resistance of
the web, if shear buckling is limiting the resistance
The web panel connects flanges and stiffeners to fairly stiff frame
corners. The maximum shear force that can be carried by the flanges
alone, can be estimated from the bending resistance of a flange. Half
the shear force is acting at the point of contraflexure, where
the bending moment in the flange is zero, see figure 5.11 b). At any of
the 4 points indicated we have:
V
Kf I w
Mf _ 2 2 2 .
v <— f bf—hfL
-£ 5.18
W{ b{hf f 3 7f’md t
6
The bending moments can as a rule not reach the full resistance of
the flange, because the shear deformation of the web panel will not
be large enough. It is, therefore, reasonable not to use more than
a little less than half of 2/ 3, say 0,3. The total shear resistance is now
the sum of equations 5.11, page 126, and 5.18. But the panel shear
strength according to the right part of equation 5.10, page 125, cannot
be exceeded. Hence we get:
%v ^z,max,Rd + ^’^/f,md
V
z,Rd = mm< w 5.19
V
z,max, Rd v,min,d
tv,Ed = Th
<f
J gl>nun,d if
gl 5.20
y,fic "gl "gl
where fgl,min,d usually is the planar (rolling) shear strength of the web
panel if not for some odd reason the shear strength of the flange
material in the longitudinal direction is smaller. Equation 5.20 is valid
if the shear stresses are assumed to be uniformly distributed along
the depth of each glue line hgl. The parameter ngl is the number of
glued surfaces per web panel and flange, that is 1 for a box-section
and 2 for an I-section. In this case all the glue lines are of the same
Tv,2,Ed
dx
GC
Section A-A
a)
Figure 5.12 Shear stresses in glued flange-to-web connections. a) Entire cross-section, b) sections 1 -1 and 2-2 cut open
indicating distribution and direction of shear stresses, c) as in (b), but for the bottom flange having a deep glue line.
pt-h
n sv
"gi_____
Sing =1 single sided glue line
Tv,red,Ed - T’ where 5.22
b Mg! = 2 double sided glue line
"gl
Figure 5.13 A possible reduction of shear stresses for plywood. a) One glued and one free surface, b) two glued surfaces.
Note that the reduction given by equation 5.22, page 131, might be
a bit too optimistic for a double sided glue line if the number of
veneers is smaller than 6.
A (' K(x)
^^shearW q A 5.23
where Gw is the mean shear modulus of the web directly from mate
rial standards, Aw is the total web area taken between flange centres
and V(x) is how the shear force varies along the beam. Direct integra
tion of equation 5.23, page 132, and determination of integration con
stants with regard to the boundary conditions result in a function for
the shear deflection. For a simply supported beam subjected to uniform
loading q, that is the most common support and loading conditions
in the world, the maximum shear deflection at midspan becomes:
qL1
23
W,shear,max 5.24
laminated timber or LVL; but web parts that are themselves compos
ite beams are not unusual.
Figure 5.15, page 134, shows one typical SSP-element, an H-section
having two thin flanges and webs of rectangular timber pieces.
Some relevant failure locations are indicated symbolically and
given a specific number in figure 5.15. The indicated failures refer to Figure 5.14 Typical cross-sections of stressed skin panels
a section with stocky web parts that are not sensitive to any kind of (SSP), H- as well as T-sections are shown
buckling. If, however, the web parts are made of thin sheeting
materials the relevant failure modes from figure 5.3, page 117, and sec
tion 5.1.1, page 116, should also be checked. The most relevant failure
modes of elements with stocky webs are:
4. Shear failure of the web parts. The resistance is verified as for struc
tural timber, for which design rules are given in section 3.1.1, page 54.
If the webs are made of sheet material, the rules in section 5.1.1.3,
page 124, apply.
5. Shear failure of flange-to-web interface, that is the glued connection
between them. The resistance is basically checked as for I-girders,
which is described in section 5.1.1.4, page 130. Note, that if the inter
face itself is strong enough, there is still a risk for a pure panel shear
failure of the flange material on either side of the web members
both for the tension and compression flanges. The exact points for
this failure mode are not indicated with numbers in figure 5.15, but
the position is almost as for point 5 but in the flanges themselves.
6. Failure of web parts with holes. No detailed design rules are given
in this book, see for example instructions from the manufacturers.
7. Flexural buckling failure caused by axial loading, not shown in fig
ure 5.15. Flexural buckling is treated as for ordinary columns in sec
tion 3.2.1, page 62.
8. Compression failure of the outer flange surfaces or the web parts in
contact with the flanges caused by concentrated forces acting per
pendicular to the grain, which is basically checked using the rules
given in section 3.1.3, page 58. But the contact area between flange
and web can be slightly increased if the sheeting material has good
compression properties in its thickness direction.
9. The influence of flange curling, that is transverse bending of both
the compression and tension flanges due to the element curvature
caused by the external bending moment. Of course some flange
curling may also be induced by transverse loading to a flange in
the regions between the webs. The basic influence of both flange
curling and transverse loading is reduced bending stiffness as
the distance between the flanges decreases. As a consequence,
the axial loading of the flanges as well as the bending deflection of
the element will increase.
10. In the serviceability limit state one have to check deflections pri
marily caused by bending. Shear deformations can at least for
stocky webs be neglected, but should be added for slender web pan
els. The influence of vibrations and resonance frequencies in floors
and bridge decks may be of profound importance. More information
about these topics is given in Chapter 7, page 183.
(Ww,max,d h-'rn.d(
(-bef (Web edge)
^.eue^e,
bf bf bf bf
bf/2 bf/2
Compression
NA NA
Tension
Figure 5.16 Effective cross-sections of stressed skin panels (SSP), H- as well as T-sections
beff /2 /2
beff If the webs are made of wood based sheeting, then the last inequality
of equation 5.25, page 135, is replaced by a verification of the edge
stresses being smaller than the tensile and compressive strengths of
the board material, as appropriate.
If glued thin webbed beams are used as webs in an SSP-element
the last inequality of equation 5.25 is replaced by equation 5.6, page 122.
Now, the methods of finding the effective cross-section with regard
to the influence of shear lag, local buckling and flange curling will
be discussed. We start by discussing shear lag in combination with
local buckling, whereafter flange curling is treated separately.
Both shear lag and local buckling cause a non-uniform stress distri
bution along the flange width. The maximum stress appears around
the web panels, while the minimum stress appears between them.
/2 Shear lag may influence both compression and tension flanges, while
/2
beff beff
local buckling is only relevant for compression flanges.
The flanges of any beam are subjected to in-plane shear stresses,
that is acting in a plane “perpendicular” to the webs. It is these shear
stresses that gradually build up the axial stresses in the flange, but at
the same time they will cause some small shear deformation of
the flange. This shear deformation will consume some of the axial
straining needed to build up full axial stress in the flange material
Buckle between the webs, resulting in a non-uniform stress distribution as
shown in figure 5.17 a). From a designers point of view this fairly com
plicated stress distribution is handled through the concept of effective
b) width. The real flange width is replaced by an effective width for
which gmean b f = gmax b eff is supposed to be valid.
Figure 5.17 Non-uniform stress distributions, in a wide
To find the best effective width in each particular case is difficult
compression flange between two webs, caused by:
a) shear lag only,
because it is influenced by many factors, out of which the most
b) local buckling only. important are: shear properties of the flange material, loading config
uration, boundary conditions, ratio of flange width to span length
and local buckling. In Eurocode 5 the effective width is simply given
as a constant fraction of the span regardless of this complexity.
The span is the distance between points of contraflexure or zero
bending moment. In figure 5.18 a comparison between the simple
Eurocode expression and theoretically more correct relations are
given for a flange made of OSB loaded either perpendicular or parallel
osb ± q
osb q
Z'
osb ± P
osb Z' P
EC5: 0,15Z
beff
bf
Figure 5.18 The influence of shear lag for an OSB flange in a simply supported SSP-element
subjected to two different loading conditions
h (h
f 5.26
°f J
where the coefficients ft0 and ft1 are given in table 5.1 together with
the traditional Eurocode expressions.
Buckling of compression flanges gives a non-uniform distribution
of axial stresses which is similar in shape to the distribution caused
by shear lag, as indicated in figure 5.17 b), page 136. Shear lag reduces
the influence of local buckling because it leads to smaller compres
sive stresses in the part of the flange that is sensitive to buckling.
The other way around local buckling will reduce the influence of
shear lag. But we start to discuss local flange buckling as a stand-alone
problem. As an instability problem it can be accounted for using
the same principal
approach as outlined for shear buckling in section 5.1.1.3, page 124,
resulting in an effective flange width (b eff = b f xc). This process may be
summarised as: critical compressive stress ^ slenderness parameter
^ reduction factor, that is using gcr ^ Xc ^ xc, where index c indicates
compression. The slenderness parameter is defined in the same way
as equation 5.14, page 127 and the design curve is chosen as in equa
tion 5.15, page 128, but index v is replaced by c to indicate compression.
Table 5.1 Effective width due to shear lag in wide compression or tension flanges
Widths according to both Eurocode 5 and equation 5.26 are given.
p 0,97 1,05
p 0,83 1,60
1) 3± denotes a three veneer plywood with the face veneers oriented perpendicular to the axial stresses.
^-0,8
%c = 5.27
where ocr is the critical buckling stress for a long plate simply sup
ported along all four edges and subjected to uniform compression in
its longitudinal direction. The factor 1,25 increases the buckling
stress by 25 % to approximately account for some of the rotational
stiffness provided by stocky web members (total rotational constraint
along the webs would give about 60 %). If, however, the web members
are thin and slender 1,25 should be replaced by 1,0, as thin board
members provide little rotational stiffness. of,c is the compressive
stress in the flange at the load level where the bending resistance of
the entire SSP-element is exhausted, under the assumption that no
local buckling affects the resistance of the flange. It is always on
the safe side to set of,c = fc,0,d or fc,90,d, as appropriate. But if the bending
resistance is governed by the tension flange or webs, rather than
the compression flange, a proportionately reduced value can be used.
Also, if shear lag is affecting the compression flange the above value
of of,c can be reduced by the factor beff,shear lag/bf to take the beneficial
effect of shear lag into account.
Note that it is not self-evident that that the design curve should be
0,8 /Ac, but it must under all circumstances have a similar shape.
A comparison to the more complicated curves used for steel in
Eurocode 3, observing that those curves also take the influence of
residual stresses into account, indicates that 0,9 /Ac might be a better
choice. But choosing 0,8 is slightly on the safe side, especially for
slender flanges, and can also be said to implicitly account for some
small influence of flange curling. This simple formula for the design
curve makes it possible to express the effective width as a constant x
flange thickness, as we shall see next.
The critical stress ocr is calculated based on linear elastic theory and
a lower limit, on the safe side, is obtained by assuming the flange to
be a long plate simply supported along all four edges. But the value
can easily be increased or decreased if other boundary conditions are
more relevant, for example increased if the flange is equipped with
closely spaced transverse stiffeners. Compression flanges in SSP-
elements are either made of isotropic or orthotropic sheet materials
and the critical stress is conveniently expressed as:
Isotropic plate
5.28
„ _k Orthotropic plate
cr hfb?
where Dx and Dy is the bending stiffness of the flange about its x- and
y-axis, respectively. Ex and Ey are the elastic moduli for bending about
the two main axes of the sheet material as if it would be homogeneous
over its thickness, that is the moduli usually given in various stand
ards and test reports for timber based sheet materials. The buckling
coefficient kaiso = 4,0 applies to long isotropic plates. For orthotropic
plates the buckling coefficient kaorth depends on two parameters:
D
and j82 =— 5.29
where G is the shear modulus. But, for long plates (l > 2bf) that are
only loaded by uniform compression in their longitudinal direction,
the value of ft1 does not matter and a fairly good approximation of
the buckling coefficient is:
JL
°’nHh
orttl
= 2,0 + 2,0 P2
& = 2,0 + JeTe 5.30
V x ;
By using kaiso = 4,0 and kGoiih from equation 5.30 in equation 5.28,
page 138, and substituting the obtained critical stresses into equa
tion 5.27, page 138, the reduction factor for the effective width is
a constant times hf/b f and the effective width itself is just a constant
x hf. Obviously the constant is a unique number for every sheet
material on the market. In table 5.2 the results of such calculations
are summarised for some common sheet materials together with
the effective widths recommended in Eurocode 5. It appears that
the precision of the Eurocode 5 constant is rather poor and too much
on the unsafe side for OSB, particleboard and fibreboard. The reason
for this difference is that the design curve is taken as 1,0 /Ac rather
than 0,8/Ac for the Eurocode 5 constant, which is definitely too opti
mistic if Ac < 1,30 and the unfavourable effect of flange curling is not
considered; or if the favourable effect of shear lag is not implicitly
taken into account.
Following Eurocode 5, the effective width should be taken as
the smallest width obtained either from the Eurocode column of
table 5.1, page 137, or table 5.2 for compression flanges and only from
table 5.1 for tension flanges. In addition, the influence of flange
curling is not even mentioned. The main advantage of this approach
is its simplicity, but its precision might be very poor. It is pointless to
perform complicated and tedious analyses of SSP-elements with
Table 5.2 Effective width due to buckling of compression flanges according to Eurocode 5 compared to values obtained
by equations 5.30 5.28 5.27
, and applied to 139 different board products for which material properties are available.
Effective width = const x hf. The two rightmost columns show the effective width (const x hw) in case of shear buckling,
as described in section 5.1.1.3. Minima and maxima obtained for different products in each category are given.
OSB, ± 25 21 23 40 45
Plywood, 3, // 20 12 15 22 27
Plywood, 3, ± 25 17 22 44 61
Plywood, // 20 17 24 35 63
Plywood, ± 25 19 32 44 73
Particleboard 30 23 26 35 40
Fibreboard 30 20 28 28 44
1) // The flange major axis of bending oriented transversely to the SSP-element, for example face veneers of plywood oriented parallel to the webs.
2) ± The flange major axis of bending oriented longitudinally along the SSP-element, for example face veneers of plywood oriented
Flange curling is a problem for thin and wide flanges and affects both
tension and compression flanges. The axial bending stresses in
the flanges will always have a very small component acting towards
the neutral axis of the SSP-element, which is a consequence of
the bending curvature. It is equivalent to a distributed load applied
to the flange surface, causing transverse bending of the flange. This
results in a reduced distance between the flanges, which in turn
reduces the second area moment and increase the axial flange stresses.
The maximum transverse bending deflection of the flange in between
two neighbouring web panels can be estimated from elementary beam
bending theory applied to a flange strip supported on a number of
webs. Following the approach taken in StBk-N5 (1980) the maximum
flange deflection can be estimated to:
5.31
where of,mean is the average axial stress in the flange calculated as (of,c,d
or o) x beff/ bf, that is the average stress with respect to shear lag and
buckling. Ef is the modulus of elasticity with respect to flange bending
about an axis parallel to the webs. Set Ef to Emean under instantaneous
conditions in service class 1, 0,8 Emean in service class 2 and 0,4 Emean for
plywood that can be used in service class 3. Set Ef to Emean/(1 + ^2 k f)
under final conditions with ^2 determined for the load causing
the largest stress. zf is the absolute value of the distance from the neu
tral axis to the centre of the flange. K is a modification factor taking
the support conditions of the flange into account and is given below.
Note that the above recipe for finding the moment resistance is appli
cable to SSP-elements for which the transverse load enters into
the element right on top of the webs. No external transverse loading
can be applied directly to the flanges in between the webs, which
would cause severe transverse bending of them. In such cases, like in
floors and outer roof surfaces, the flange must first be designed such
that it can carry this transverse bending from distributed and/or con
centrated forces. After that the recipe can be applied, but the flange
slenderness is then usually so small that only shear lag needs
attention.
Figure 5.20 Three degrees of composite action, for equal midspan deflections
Core
Original
configuration
Displaced configuration
x
Figure 5.21 The beam cross-section and a small cut-out d along the beam axis
In the figure the following notations are used (other symbols are explained as they are introduced):
The directions of the arrows in figure 5.21, page 144, indicate positive
forces, moments and displacements. The prime symbol, ', indicates
differentiation of the above functions with respect to x, two prime
symbols indicate that the differentiation is performed twice, etc.
M = «2 - M1 + Atw' 5.32
where:
du.
£. = —1
1 dx
du.
E =--- —
2 dx
are the simple engineering strains along layer centres 1 and 2, respec
tively. We have also assumed both layers to have the same deflection
w(x) and approximately the same radius of curvature. The strains
may, by using traditional bar theory and positive tensile strains, be
expressed as:
N2
£.2 =-- — + £tt2, 5.35
where st1 and st2 are uniform strains caused by temperature changes
in layers 1 and 2, respectively. By substituting equations 5.34 and 5.35
into equation 5.33 we obtain:
Now, since both st1 and st2 are constant along the beam the difference
between them is replaced by one constant:
Force equilibrium for the cross-section shown in figure 5.21, page 144,
requires that:
N3 = -TVj 5.38
m' = Aw"-
fl 1
------ +-------- ^+£t 5.39
‘ 1^4 ea^ 1 ‘
Let now the material properties of the core material enter into this
derivation. When the shear flow H affects the infinitesimal cut-out in
figure 5.21, page 144, it is displaced an angle that may be written as:
Shear stress
y =----------------- H
=— 5.40
Shear modulus G b
H -N'
5.41
Gb
where the shear stiffness of the core, kcore, has been introduced:
* ""
core i
“core
1
kcore
= Atw"- A + £i 5.43
M4 M,+M,-.X,h. 5.44
,.__A£ m2
5.45
£A ei2
5.46
EIy + EI2 EIr + EI2
1 l 1 t A2 ' k h
^core
N. = —5-^r M - k et 5.47
1 E A A A+A> 1 E (Zj + ^J core ‘
q _
kcore h
5.48
■E
2 k ( 1 1 A2
£02 = _c»e _L + _L +---- 1--- 5.49
e A A A+A,
1 1 = CM-kcore ett
N'-o^N. 5.50
5.51
where:
5.52
5.53
f n" C k e
M+h 5.54
jo2 co2 co2
5.55
5.56
-1 A! + A2
-1 1-1
E J +1 + ^fca
1 2 4 + ^2
5.57
is the second area moment of inertia for the cross-section when full
composite action prevails. Using the result from equation 5.56 in equa
tion 5.55, page 147, gives:
„ -M C
w =------------------ 5.58
EI{ica core
Finally, by integrating equation 5.58 twice, the link between the deflec
tion function and the axial force function may be expressed as:
M.------ 5.60
1 I'+I^ 1 t7
i=0
( k E
N'.'~(D
1 2N.=C
1 n —0016 1 + a.x + a~x2 + o,x3 + a.x4
a0/^1234 5.63
5.64
’ i=0 ’ i=2
^core£t c 2
+ xax + x2
C
ao
c I ®2> ‘2
5.67
co2 6xy 12x2 24
+ 3
+
»2> o4
+ DyX + —
c 5.68
^core®2
(o£ CP
= Z>2®cosh ft)2 2 ° d2 =
ft)l
2 G)3 cosh 5.69
~T
The total solution is obtained by substituting the results of equa
tion 5.69 back into equations 5.67 and 5.68, and after some rearranging
of terms and developing of the deflection function (wfca for full com
posite action) we arrive at:
sinh(fi)x)
-(OX
, f tot'
cosh —
I 2 ,
lx
5.70
8ELtea 2
sinh(<ax
c2 ------- 7---- x— U'-A
J 1
< cosh I — 2 J ,
One can of course find solutions to other problems, being loaded dif
ferently and having other boundary conditions, by using the same
principal approach. A summarising recipe may look like:
eee fca
------ PCA
••• NCA
x-coordinate / m
Figure 5.22 Partial composite action in comparison to the upper and lower limits
, kn
k =— 5.72
s
where k is the slip modulus of a single connector (for instance Kser or
Ku for a timber structure designed according to Eurocode 5), s is
the constant spacing between the connectors in the longitudinal
direction and n is the number of connectors placed on a line perpen
dicular to the longitudinal direction. Suppose that two (or more) dif
ferent types of connections connect the two layers, for example a shear
flexible core of thickness hcore and mechanical connectors with slip
modulus k and spacing s. One must then distinguish between parallel
or series arrangement of the different connections. In the parallel
case both types of connectors connect the two layers independently
of each other, for example mechanical connectors that go right
through a core material. In the series case the shear flow is trans
ferred from one of the layers into the first type of connection then
into the next, etc., until the second layer is reached. An example is
a shear flexible core that is connected to each layer by mechanical
connectors that do not fully penetrate the core material, but have
the same properties at each side of the core. For the discussed exam
ples the stiffness in the parallel case may be written as:
, kn Gb
k<°
-~
* +r~ 5.73
core
and for the series case it may be expressed as:
1
5.74
2 S ^core
kn Gb
Equations 5.73 and 5.74 are merely examples and must be properly
adjusted in each particular case. For more complicated interlayer
connections, as in latticed girders and girders with packs or gussets
(Vierendeel type), expressions for kcore are for instance given in Larsen
and Riberholt (2005).
In connection to equations 5.40 to 5.43, page 146, is explained that
the shear flow along the interface is the first derivative of the axial
force in layer 1, i.e. H = -N'1. The shear force acting on an individual
connector is therefore:
£
57x) = -N!(x)— or 57x) = ku(x) 5.75
n
and from equation 5.41, page 146, the slip along the interface is
expressed as:
f \ • | 7£X I
?w=^osm — ; M" = -q gives 5.77
Tlx
sm 5.79
~T
By substituting equation 5.79 into equation 5.59, page 148, noting that
D3 = D4 = st = 0 and topping off with some algebraic manipulations
the deflection function is found to be:
4 1 +------ -
?o^4
w(x) = /dmfSin 5.80
^4^4a 1+P Y
+ l at J
C_kcore *t
valid for DS3L
A1, I1
Undeformed configuration
h1
h2
h1 = h3
A3 = A1
I3 = I1
E3 = E1
N1= - N3
The more general single symmetric three layered case is more com
plicated. A detailed derivation will result in a system of differential
equations, where each equation is similar to the one given in equa
tion 5.50, page 146. In general one equation is needed for each inter
face between the parts, that is two equations for a 3 layered beam,
three equations for a 4 layered beam, etc. Equations for the 3 layered
case can for instance be found in Oduyemi and Wright (1991).
Table 5.3 Summary of equations for 2 layered and 3 layered double symmetric beams. No st-term is present.
Interface slip
m(x) = ~~rL— or M(x)= “-^1 (x) A— = —1—
£core k
Relation between c 2C
N1 and w w-wfca , 2Ni + D3x + D4 W-Wfca , 2-V1+Z>3X + D4
Acre®
Constants k A k A
Q— core Q— core
’E ( A + A 2E | 2r,+r,
(1 f
n _ kcore
—+ ±+_?L' g,2 = Arae 1
U + 2 (2Z1+A)J
*,2 1
E ■^2 A+A j E
c i 1
G+A, 2Z1 + A)
k r i _
k
con
(J _ core
M E IV M EA
"A,
Bending moment in
M.------ M + N,ht) M. = M,-------—(M + N. h\
parts 1 and 2 1 W 1 3 2^ + 1^ 1
Equations 5.72 to 5.74, page 153, give possible relations between kcore
and k for mechanically connected parts or layers. Equations 5.70,
page 151, 5.71, page 152, and equations 5.79 to 5.81, page 154, give solu
tions to 3 different elementary cases of practical interest.
An increasing slip modulus will give the opposite effects to those listed
above. Luckily a large change in the slip modulus will produce a much
smaller change in member forces, member stresses and deflection.
Reducing k to half its value or doubling it will change these values in
the order of 10 %, under typical conditions. But note that the connec
tor force will typically change by the same order as k.
From the above description it is recommended that Ku is used when
checking resistances in the ultimate limit state (ULS). This holds even
if the weakest part of the beam is the connectors, because then their
slip modulus will be even smaller than Ku for the most loaded (and
Table 5.4 Recommended values for the slip and elastic modulus of composite timber beams
for which mechanical connectors are used and partial composite action prevails
^ULS.inst = *
u
k________
1 + 2^
* A note on determination of final deformations. Cross-sectional constants, that is Ifca, C, rn, CMb and CM of transformed cross-sections,
should be based on ESLS,fin and kSLS,fin for the instantaneous condition before the effect of creep is added. But all moduli beyond these
fictitious cross-sectional constants should keep their instantaneous values ESLS,inst = E0,mean and kSLS,inst = Kser, because they will at a later
stage be multiplied by either (1 + kdef) or (1 + 2 kdef) when going from instantaneous to final deformation as given by equation 5.86.
Wfca,SLS,inst + Wslip,SLS,inst+
w
vinst = (Qi) +
Wfca,SLS,inst (^1) + ^sUp.SLS.inst 5.85
Wfca,SLS,fin (fil )(^ + V^2,l ^dcf ) + Wslip,SLS,fin (fil )(^ + ^2,1 ^dcf ) +
5.86
X(Wfca,SLS,fin(fii)(v<O,i + ^2.i^def )+ Wslip^LS,fin (ei)(ni+2nifcdrf))
y
where 2 is a reduction factor giving the quasi permanent load value,
y
while 0 gives the combination value of the load.
In equations 5.85 and 5.86, make sure to use E0,mean in all wfca and Kser
in all wslip. Note also that shear deflection caused by shearing of each
timber member is neglected, which is justified by the fact that its
contribution usually is small compared to wslip.
Horizontal stabilization
Roberto Crocetti
163
6.1 General considerations
6.3 Fundamental statics 164 In design of structural systems, the way lateral stability is achieved is
of fundamental importance. For the sake of simplicity let us
6.4 Strength and stiffness requirements approach the problem by studying a planar structure. All structures
for bracing systems 167 undergo some changes in shape under load. In a stable structure,
6.4.1 Perfectly straight column 167
the deformations induced by the load are typically small. Moreover,
6.4.2 Column with initial imperfections 169
6.4.3 Beam, truss or column systems 171
in a stable structure, internal forces are generated by the action of
the load; these forces tend to restore the structure to its original
6.5 Typical bracing systems shape after the load has been removed. In an unstable structure,
for heavy timber structures 174 the deformations induced by a load are typically large and tend to
6.6 Special topics 176
continue to increase as long as the load is applied. By contrast,
6.6.1 Forces arising due to an unstable structure does not generate internal forces that tend to
the slope of the roof 176 restore the structure to its original configuration. It is the fundamen
6.6.2 Stability of the upper chord of tal responsibility of the structural engineer to assure that a proposed
low-arch or low-truss bridges 177 structure does indeed form a stable configuration. Stability is a cru
6.6.3 Bracing of continuous beams cial issue in the design of structures that are assemblies of discrete
at intermediary supports 177
elements. For example, the post-and-beam structure illustrated in fig
6.7 Bracing of timber framed buildings 178 ure 6.1 a) is apparently stable. However, any horizontal force tends to
6.7.1 Floor and roof diaphragms 178 cause deformations of the type indicated in figure 6.1 b).
6.7.2 The in-plane resistance of shear walls Clearly, the structure has no capacity to resist horizontal loads, nor
under horizontal loading 179 has it any mechanism that tends to restore it to its initial shape after
horizontal load is removed. The large changes in angle that occur
between members characterise an unstable structure that is begin
ning to collapse. Consequently, this particular pattern of members is
referred to as a collapse mechanism.
There are just a few fundamental ways of converting a self-stand
ing structure of the general type shown in figure 6.1 from an unstable
to a stable configuration. These methods are illustrated in figure 6.2,
page 161.
The first is to add a diagonal member to the structure, see figure 6.2 a). Table 6.1
In such a way, the structure cannot undergo the “parallelogram- Diameter Horizontal stiffness
ming” indicated in figure 6.1 a), page 160, without a very large elonga k F/
= d
tion of the diagonal member — which will not occur if the diagonal
No diagonal k = k0
is adequately sized. Another method used to assure stability is
d = 10 mm k » 7 • k0
through shear walls, that is rigid planar surface elements that resist
changes in shape of the type illustrated, see figure 6.2 b). A third d = 20 mm k » 25 • k0
6m
or the other of these basic approaches for stability. More than one
approach can be used in a structure, for example a structure having
both rigid joints and diagonals. This is normally done when the stiff
ness of the structure needs to be improved and, in such a case,
the redundancy is obviously increased. As an example, consider
the following glulam frame, with clamped connection at the base of
the columns and pinned connections between beam and columns, see
figure 6.3. The system is stable without any diagonal bracing. However, Figure 6.3 Influence of the diagonal bracing
the introduction of a steel diagonal dramatically increases the hori on the stiffness of the frame
zontal stiffness of the frame, even when relatively small diameters of The column are bent around their strong axes.
Reality Codes
Figure 6.5 Wind pressure and suction, reality versus building codes. In some building codes the model for
wind pressure and suction may be more sophisticated with slightly different coefficients.
Wind forces on the side of simple buildings of the type shown in fig
ure 6.5, for example, would be picked up by surface members, which
in turn transfer them to secondary frame elements. Earthquake forces
have a slightly different character than wind forces, as they are nor
mally concentrated at high-mass areas, such as roofs or floors, but for
the purpose of this discussion, we assume that the effects of wind or
earthquakes can be represented by a series of forces acting laterally
on the primary structure.
It should be emphasised that a primary role played by a roof or
a roof plane is, besides resisting vertical loads, to transfer lateral
forces to side shear walls, cross braces, or frames. Figure 6.6, page 163,
illustrates two methods for picking up forces acting on transverse
wall and transferring them to side shear planes. The figure also illus
trates how forces on the longitudinal face are transferred to side
reverse shear planes.
The use of an edge beam approach illustrated in figure 6.6 a) demands
that the edge beam be sized for both vertical and horizontal loads.
This approach is often used for relatively small buildings. As building
dimensions increase, a larger part of the roof plane is typically uti
lised for stiffness. An entire bay, for example, might be cross braced
as a way of providing sufficient in-plane stiffness for a large building,
see figure 6.6 b). In these cases, it is essential that the whole system be
organised such that the rigid horizontal plane (at the roof) both receive
externally induced horizontal forces and transmit them to side shear
planes (at the walls). Unless there are two separate systems for trans
mitting horizontal forces, for example purlins and struts, the floor or
roof beams (purlins) must be designed to carry both bending due to
vertical loads and compression due to horizontal loads.
Purlin
Loading on gables
a) Forces from secondary frames are resisted by an edge beam b) Forces are transferred through roof members to the roof
with high lateral strength and stiffness, and carried directly to diaphragm which transfers loads to side diaphragms (note that
side shear walls or diaphragms (typically small structures only). in this case purlins must be designed to carry both compressive
forces and bending from vertical load).
Figure 6.6 Force transfer to rigid wall or roof-planes. The graphic convention of a grey tone with dotted “X”
is used herein to depict a rigid plane, regardless of whether it is a wall, truss, or rigid frame.
When steel rods are used, they are usually arranged in crossing pairs,
so that if racking puts one of the elements into compression (which it
is incapable of resisting), its complement will resist racking by tension.
Bracing may also be necessary to prevent lateral buckling of
the entire roof plane and to keep the filler beams (purlins or separate
struts) in place so that they can give the proper lateral support to
the compression zone of the beam under in-plane bending, see fig
ure 6.7, page 164.
Buckling length: 0
Figure 6.7 Roof structures with main beams braced in different ways:
a) Unbraced. Buckling length of the main beams lef = l.
b) Braced with crossing pairs of steel rods. Buckling length of the main beams lef = l/4.
c) Braced with profiled metal sheathing. Buckling length of the main beams lef = 0.
The wall bracing system which resists lateral forces may be arranged
in a symmetrical or asymmetrical manner. For the symmetrical case
they will only carry direct forces as caused by translation. For
the asymmetrical case, additional forces due to rotation (torsion) are
generated, see figure 6.10 b), page 165.
6.2
V = H-^ = ^ 6.3
1 3-R 6
Roof with wood based sheathing, or with some type of steel deck, are
generally considered to be flexible diaphragms. They are normally so
much more flexible in comparison to the lateral supporting system
that the supporting structures are considered to be rigid. The behav
iour of the flexible diaphragm may be visualised as simply supported,
discontinuous beams sitting on unyielding supports. Thus, the loads
can approximately be distributed as based on tributary area. For
the case in figure 6.9 a), note that the smaller interior walls carry dou
ble the force as the longer exterior walls, because the tributary areas
are twice as large as the ones for the exterior walls; the stiffness of
the walls is not considered since they are assumed infinitely rigid.
The behaviour of semi-flexible or semi-rigid diaphragms, on
the other hand, is highly indeterminate. They are assumed to distrib
ute the loads like a continuous beam on flexible supports.
If the lateral force-resisting structures are arranged in an asymmet
rical fashion, then in addition to translation (that is direct shear),
rotation (that is torsion) is generated. The magnitude of the generated Figure 6.9 a) Flexible roof diaphragm,
rotation is also influenced by the building proportions, see figure 6.10. b) semi-rigid roof diaphragm.
The solution for the distribution of the lateral forces to the asym
metrically arranged shear walls can be approached by distinguishing
between determinate and indeterminate systems as based on the fol
lowing simplified assumptions:
• Consider the roof as a rigid diaphragm. Flexible diaphragms are
considered incapable of transmitting rotational forces.
• Neglect the bending stiffness of the walls about their weak axis.
• Consider intersecting walls as separate units. There is no shear
transfer at the corners.
• Consider the shear flow along the wall to be constant.
• Neglect the effect of columns in resisting lateral forces.
a)
than three shear walls are present, the structure is statically indeter
minate and the force distributions can be found by means of either
approximate methods or computer-based approaches. As an example,
let us study the force distributions in the building shown in figure 6.11.
The building has a channel-shaped stabilising system and the hori
zontal loads act mainly along two orthogonal directions, that is x and
y. Let us first consider the equilibrium for loads acting along y-direc-
tion, see figure 6.12.
The total horizontal force “Y” acting perpendicular to the long
fagade of the building must be resisted by the forces P1, P2 and P3. It is
worth to remember that the wall assembly is not considered as a single
channel-shaped unit but as three individual, independent elements.
Moment equilibrium about the intersection of the unknown forces P1
Figure 6.11 Illustration of a building stabilised by and P3 yields the unknown wall force P2:
a channel-shaped bracing system
xn 2
^M = 0 => y-2 a-P2-3 a = 0 => P^ = ~ Y 6-4
In fact:
Figure 6.12 Resisting mechanism for horizontal loading Figure 6.13 Resisting mechanism for horizontal loading
perpendicular to the long fagade perpendicular to the short fagade
Ca „ 2P
Unstable Manst > Mst => P>----- =} C <----- 6.11
2 a
,, .„ _ Ca „ 2P
Stable *
Afu^cM => P < — => C >----- 6.12
2 a
Ca 2P
Critical na = M.sit => P =----- => C =-----
M„unsi 6.13
2 a
If the equilibrium is stable, it means that when an external perturba
tion is applied, and when the source of this perturbation is removed
the system will return back to its original position, that is the perfectly
straight position. On the other hand, if we imagine to gradually
increase P, and for each small increment of P we give a small pertur
bation at the mid-height of the column that produces a displacement
A, a condition will be reached in which the column — after the given
perturbation — will not return to its original position. The load cor
responding to this situation is referred to as the critical load.
However, the maximum axial load that the column can resist cannot
be larger than the buckling load of the column, which in this case is:
, E l
P<PE = n2- — 6.14
a
PE is the Euler load for a strut with the length a under compression, E
is the modulus of elasticity parallel to grain and I is the second
moment of area about the z-axis, according to figure 6.14, page 167.
The corresponding buckling shape, when the stiffness of the spring
C > CE is large enough for achieving either a stable or a critical equilibrium,
is shown in figure 6.15.
The spring stiffness CE which is necessary in order to achieve
the buckling shape shown in figure 6.15 is therefore:
„ 2-Pp „ 2 El
CE------ - = 2-n2—3- 6.15
P a a
Figure 6.15 Buckling mode for spring stiffness C> CE Figure 6.16 shows the relationship between the load carrying capacity
of the column and the stiffness of the bracing element, for an ideal
column. When C < CE, there are two equilibrium situations, one stable
PE
with no transversal displacements and one critical with a small trans
versal displacement. This critical situation is achieved when the rela
tionship between P and C is that given by equation 6.13, page 166, (in
figure 6.16 this relationship is described by the line AB). When C > CE
there will also be two possible equilibrium situations, again, one stable
PE with no transversal displacement and one critical when P = PE, with
the formation of a buckling shape with two half-waves, see figure 6.15.
The above model — though slightly more complicated — can also
be used to determine the critical load on a column with several inter
mediate braces each with the same stiffness C. Figure 6.17 shows
a column with three intermediate equidistant braces. Increasing
the brace stiffness, the buckling mode of the column switches from
PE a single half-sine wave to two, three and finally four half-sine waves,
4 which occurs when C approaches the value (3,4 • PE ) / a, Yura (1996).
If the number of braces is further increased, the brace stiffness nec
CE essary to form a series of half-sine waves between all the brace points
Figure 6.16 Influence of the brace stiffness on
is slightly larger. For practical cases, when the number of intermedi
the loading carrying capacity of an ideal column ate braces is n > 4, the required stiffness should be CE > (4,0 • PE)/ a,
Note that PE and CE are defined in equation 6.14 and Yura (1996).
equation 6.15 respectively.
T 2
Let us first study the system when the stiffness of the spring is C = CE,
that is stiffness such to make the column to buckle according to fig
ure 6.15. We impose:
2-P■tif
C = CE = 6.18
a
Substituting equation 6.18 into equation 6.17 gives:
For C = CE and A0 ^ 0, the heavy solid line in figure 6.19 shows the rela
tionship between P and AT/A0 given by equation 6.19. For P = 0, AT = A0.
When P increases and approaches the buckling load PE = n2 • E • I/ a2,
the total deflection AT becomes very large (theoretically AT tends to
infinity). For example, when the applied load P is about 90 — 95 % of
PE, the total deformation AT becomes as large as 10 — 20 times A0.
Figure 6.19 Relationship between Pand AT for a braced column with initial out-of-straightness
In this case, the deflections are much smaller than for the case with C
= CE, as shown in figure 6.19. For example, when the load just reaches
the buckling load (P = PE), then AT = 2 • A0, not infinity as in the case
with C = CE. For C = 3 • CE and P = PE, AT = 1,5 • A0.
It is evident that the larger the brace stiffness, the smaller AT. Let
us now estimate the forces that occur in the brace when brace stiff
ness is varying (note that C is derived from equation 6.17, page 169):
2P 4) + ^ 2P
■A-------- A. 6.21
a A J a
We study first the system when the stiffness of the spring is C = CE. In
that case (see also equation 6.19, page 169):
which shows that the brace force is directly related to the magnitude
of the initial imperfection. If a member is fairly straight, the brace
forces will be small. Conversely, members with large initial out-of
straightness will require stronger braces. Equation 6.22 can be rewrit
ten in the following manner:
P_ = 1___ L_ LA 6.23
Pe Fbr/P «
A plot of equation 6.23 for an initial imperfection A0 = (2 • a) / 500 is
shown in figure 6.20. If the brace stiffness is just equal to the ideal
stiffness, then the brace force becomes very large as the buckling
load is approached; this occurs because AT becomes very large, as
shown in figure 6.20. For example, at P = 0,84 • PE, equation 6.22 gives
Fbr/P
P F
Figure 6.20 Relationship between and br for a braced column with initial
A
out-of-straightness equal to 0/(2 • a) = 1/500
N
6.4.3 Beam, truss or column systems
Where a bracing system is required to provide lateral stability to
a series of compression or bending members (for example columns, N
KT A 8-N-A0
N ’ Ao - <7hr g => tfhi - ^2 6.24
8-jV-J0
?h =w' 6.25
£2
Sine-shaped
N N N deformed profile
Bracing system
Figure 6.22 Model for the determination of the brace forces: Figure 6.23 Equivalent lateral load acting on the bracing system
bracing system for a series of compression or bending members
These results show that the brace force increases with decreasing lat
eral stiffness of the bracing system.
Experimental results also show that in case of double pitched
beams with upper side inclination of 5°, the brace force is 10 % to
30 % higher than in the case of similar timber beams with parallel
upper and lower sides, Holzbau Handbuch (2000).
B = 25 mm, H = 300 mm
—o— Bracing beam: UPE 50, I = 9,12 x 106 mm4
—•— Bracing beam: UPE 80, I = 19,4 x 106 mm4
G = 500 MPa
E = 11 000 MPa
Figure 6.24 Test setup for determining brace force, and test results
The struts are connected to the top edge of the tested timber beam (Holzbau Handbuch, 2000).
9h n'30h-Ar 6.28
where:
M maximum moment in the beam.
h depth of beam.
l span of the beam [m].
n total number of laterally braced beams.
kcrit factor that takes into account the risk for lateral buckling
in an unbraced beam (see section 3.2.3.3, page 68).
*2 “ ”b 6.29
n
where:
a distance between purlins or, in sheets direct on
the roof beams, distance between screws.
nb number of roof beams braced by the force transferred
via the purlin or roof sheeting.
The bracing structure shall be stiff enough to limit the deflection due
to the design load qh to l / 700 and due to the total load, including for
example wind load, to l / 500.
] <— qw
A-A
qw
n beams
Figure 6.25 Lateral bracing of roof beams. 1) Laterally bracing structure, 2) roof beams.
a) The diaphragm action in flat profiled sheet metal roofs can be used
to stabilise glulam timber structures under the following conditions:
Figure 6.27 Examples of structures which are stable in
• The sheets forming the roof can be jointed so that shear forces can
their own plane
be transferred.
• The roof diaphragm is stiffened along its edges by beams designed
and fixed to take up axial forces. Edge beams at right angles to
the profiling can be of timber, while edge beams parallel with
the profiling are most easily made in steel.
The roof beams (primary beams) are often used as bottom and top
chords in the truss, with some of the purlins as posts (or compression
struts). The purlins, if they are used as posts for the wind truss, can
not be designed as Gerber beams. Only diagonals are added as stiffen
ers, often crosses of steel rods or sometimes timber diagonals, which
are stiff in compression. The joints between the various components
forming part of the wind truss must be designed with regard to
the forces which arise and to eccentricity. Truss posts and chords are
often in different planes.
The wind truss is often designed so that it can be used for stabilisa
tion during the erection of the structure. In smaller buildings (say
with a longitudinal dimension of about 30 — 40 m), only one wind
truss in the roof can be sufficient. The purlins or the roof diaphragm
must then be able to transfer both compression and tension forces
from the other end. In longer buildings it can be suitable, on the other
hand (not least with regard to stability during erection) to arrange
two or more wind trusses in different bays within the building.
The wind truss can with advantage be placed in the second bay from
the end. The design of joints is then not affected by the end wall
structure (gable), which usually differs from that in the rest of
the building. Other types of bracing arrangements for achieving dia
phragm action in the roof are shown in figure 6.29.
Since roof diaphragms or wind trusses are often used as lateral
bracing of roof beams, they must — in addition to other horizontal
loads — be designed for a horizontal, uniformly distributed load q,
derived according to the method shown in section 6.4.3, page 171.
The structure in hall-type buildings with large spans (> 15 m) contain
ing spaces where large numbers of people are present at the same time,
for example sports or exhibition halls and department stores, shall be
designed in such a way that the risk of progressive collapse as a result
of an accidental load is acceptably small.
It is important to observe that heavy timber structures must be
able to transmit loads from the roof level onto the foundations.
Moreover, stability must be ensured against all types of mechanisms
that can occur, including lateral buckling. figure 6.30 illustrates this
concept. If a bay of a structure is to be stabilized against lateral loads,
the first thing to do is to prevent columns to move laterally which
may be achieved, for example, by introducing diagonal members as
shown in figure 6.30 a). However, although the columns are now
Figure 6.30 Example of a proper lateral bracing of a bay in a larger timber building
braced, loads from the roof plane cannot be transmitted to the foun
dation through diagonal members. Therefore, the beams need — at
their support points — a bracing system which permits the trans
mission of loads from the roof to the underlying structure, for exam
ple as shown in figure 6.30 b), page 175. Although the system may look
stable now, the beams are still prone to lateral buckling, which may
be prevented for example by creating a horizontal truss at the roof
plane, where the beams work as compression and tension chords, see
figure 6.30 c), page 175.
An accidental load can for example be a car which runs into a col
umn, or a gas explosion within the building and as a rule it is not
reasonable to design individual structural components to withstand
such loads. Instead, the structural system should be designed so that
a primary part of the building, such as a main beam or column, may
collapse without the building as a whole collapsing. In loading com
binations with accidental loads, other loads are as a rule assumed to
act with their normal (frequent) values and not with their extreme
values (characteristic values).
Normally, it is sufficient to estimate the horizontal stability of
the remainder of the building if one main beam fails. In buildings
containing spaces for large numbers of people it must also be shown
that the damaged bay can be spanned by an alternative load-bearing
system, for example by utilising the line- and membrane effect of
purlins and roofing. This can however also be a disadvantage in cases
where a primary girder fails due to a design or construction error
which is similar for all girders in the building. In that case adjacent
girders may also fail in a progressive manner following the failure of
one bay. Several such cases have been experienced during recent
years; see Sorensen et al. (2011). The risk of progressive failure is then
higher if the bays are structurally connected to each other in the lon
gitudinal direction of the building. In such situations a better strategy
is to decouple the system to ensure that failure in one bay will not
q spread to adjacent bays.
In buildings, for example of hall type, the primary beams are often
made continuous, that is with the beam supported by more than two
columns. As a rule, the horizontal frame bracing, for example wind
truss or roof slab, is at the level of the top of the beam. Therefore, in
the zone around the intermediate support, the beam — which is sub
jected to compression at the underside — must be braced at the under
side in order to prevent it from buckling sideways (lateral torsional
buckling). The bracing of the underside of the beam can be performed
for example as shown in figure 6.33.
The bracing should preferably be such that it can only take ten
sion. This is to avoid possible torsion of the beam due to uneven load
distribution which could occur, for example, when strong wind takes
place after snowing. In such a case big masses of snow could collect
in one span of the beam leaving the adjacent span more or less Figure 6.33 Bracing of the underside of a continuous
unloaded. Thus, if the bracing diagonals shown in figure 6.33 were beam at intermediate supports
able to take both compression and tension, the beam would twist on
the support with possible risk for collapse of the roof or part of it.
The horizontal force acting on the bracing can be calculated by
the following formula, Limtrahandbok Del 2 (2016):
6.30
70 h
where:
M the moment in the beam at support.
h the depth of the beam.
Diaphragm
Edge beam
Main beam
Figure 6.34 Diaphragm action of the roof (flexible diaphragm, for example
wood based panel sheathing)
mdA =------------
g 6.31
Hence, the boundary members must carry the following axial com
pression and tension forces, respectively:
„ „ M w-f2
T = C--------------- 6.32
b 8 b
All of the bending is assumed to be taken by the edge beams and con
sequently they must be continuous or detailed to be able to transfer
the tensile or compression loading to adjacent sections.
T
M
v
c R V
c
c Plate girder
model analogy
v
v
I
Figure 6.35 “Plate girder” analogy for a roof structure: model and assumptions
The roof skin transfers the shear to the vertical braced frames (shear
walls, diagonal bracing or rigid frames). The maximum shear the roof
web must carry is:
6.33
All of the shear must be taken by the panel material. The shear stress is
higher closer to the edges of the diaphragm. The shear flow, v, (N/mm)
which the diaphragm and its connections must be designed for is:
V
v=—
b
This means that the force in individual fasteners connecting
the sheathing to the frame must be designed for a force Fv = v • s,
where s is the spacing between the fasteners. Similarly the sheeting
panel must be designed for an in-plane shear stress t = v/ t, where t is
the thickness of the panel.
h
Figure 6.37 Wall diaphragm resisting racking loads: a) Unloaded wall diaphragm, b) racking resistance, c) overturning.
HA = f\,Ab 6.35
„ b
Hd=-Fvd 6.36
s
H3=£Hdj 6.38
i
Unlike strength considerations it is very rare that serviceability con 7.1 Performance requirements 183
siderations, such as functionality and appearance as well as comfort
7.2 Static deformation 185
of the users, alone can lead to the collapse of a component or struc
7.2.1 Deformation 185
ture but they are nevertheless very important in design for a number
7.2.2 Long term and climate effects 187
of reasons, for example: 7.2.3 Load combinations 189
• to keep the visual appearance and functional requirements of 7.2.4 Limitation of deformations 190
the component or structure within acceptable limits 7.2.5 Methods to avoid deformation problems 192
• to prevent damage to brittle finishes
• to prevent undue deflection of roof structures so that for instance 7.3 Vibration 193
7.3.1 Dynamic loads 194
rainwater ponds can cause leakage through the structure
7.3.2 Response to dynamic loads 195
• to limit the effects of creep 7.3.3 Design approaches to limit
• to provide sufficient stiffness in the construction so that vibrations vibrations in timber floors 198
do not lead to problems, for example with comfort. 7.3.4 Avoidance of vibration problems 201
Some of the notations used in this chapter coincide with the ones
adopted in Eurocode but the discussion is general and describes
the phenomenon and the design methods that can be used in service
ability state analysis.
The designer may then refer each specific situation to one of these
categories.
net, fin
the effect of the shear forces, and the significance of the shear defor
mation will primarily be a function of the ratio of the modulus of
elasticity E of the member to its shear modulus G and the depth to span
ratio, Porteous and Kermani (2007). Consider, for example, a simply
supported rectangular beam of depth h and design span l carrying
a point load at mid-span. The ratio of the instantaneous deflection at
mid-span caused by the shear forces, winst,v, to the instantaneous
deflection at mid-span caused by the bending moment, winst,m, will be:
Winst,v J ; £ f
7.1
Wmstjn ’ G UJ
A-^
7.2
E,
7.3
where b is the ground plate width, l is the length of the loaded surface
and P is the applied load. Experiments have shown that in the case
where the load does not cover the entire surface the deformation is
less than that defined by equations 7.2 and 7.3.
The shorter the bearing length, the less deformation is obtained for
maintained stress level, which can be seen as an increase in stiffness
with decreasing bearing length. Another factor that affects the behav
iour is the influence of different ground plate depths, d. Directly under
the surface of the loading plate (or stud) the stress will be equal to
the value given by equation 7.3. If it is assumed that the ground plate
is resting on the whole bottom surface, the stress would be less at
that boundary than at the top boundary, since the loaded area is
Figure 7.4 Typical layout of a bearing joint larger there. This would mean that the deformation over the ground
-------- LVL
-------- Glulam
-------- Structural timber
-------- I-beam
Figure 7.7 Deformation of a joint where timber is subjected to constant compression perpendicular to grain. Same stress level 2,6 MPa in all cases.
The joint had been conditioned to 60 % RH before loading while it was exposed to RH = 30 % during the whole loading period, Martensson (2003).
Characteristic combination:
ZGkj+ek,i+X^oAj 7.9
ya ;>i
Frequent combination:
£Gkj+nAi+5>2A,i 7.10
JU i>l
where y1 1 Qk1 is the frequent value of the load Qk1 and y2i Qki is
the quasi-permanent value of the variable load Qk,i. This is the combi
nation to be used for the assessment of effects that are reversible,
that is they occur with a certain frequency but the effects will be
reduced when the load decreases again.
Quasi-permanent combination:
7-11
j>l i>l
The final result for permanent load and n variable actions will be:
n
Wfln = WfinjG + ^fin.Q, 7-15
i=2
It is important to notice that this load combination will give high val
ues for the total deformation since the deformation from the variable
load is calculated based on the characteristic value. This is reasonable
if the control of deflection concerns permanent damage, since
the effect of the high loads occurring for short time periods are then
of interest.
In the case of control with respect to appearance and utility it can
be more suitable to calculate the deflection based on the frequent
load combination equation 7.10, page 189, or the quasi-permanent load
combination equation 7.11. The formulas 7.13 and 7.14 for calculating
the final deformation from variable loads will then be modified.
Using for instance the frequent load combination as a basis we get:
7.3 Vibration
Serviceability requires consideration of the comfort of the user, and
into this comes dynamic effects such as vibration when persons walk
across a floor. This is a complex topic since it relates to the mass of
the floor structure and the actual arrangement of structural members
in the floor providing lateral distribution of applied dynamic load
and damping of the generated vibration.
Users of buildings and other engineered structures sense low-fre
quency motion in three ways:
• acceleration causes forces on the body that are felt by the balance
organs
• visual cues (for example movement of objects resting on or hang
ing from the structure relative to the observer).
• audio cues (for example creaking or rattling created by motion of
the structure).
Surveys for domestic buildings indicate that audio cues are very
important to perceptions of how well floors perform, Smith (2003).
Human induced vibrations in structures are almost always a prob
lem of serviceability, in that they are a source of annoyance to
the users. In some instances the person experiencing the motion is
also the cause of it, while in other instances it is activities of others
that cause annoyance. Thus, the activity of the person experiencing
the vibration is important. When a person walks across a floor he or
she will tolerate much larger amplitude vibrations than when sitting
quietly resting, reading or writing. Categorisation of human percep
tion, and tolerance, needs to reflect both the activity being under
taken and the relationship between the source and the sensor. In this
respect, the following definitions are used (Smith, 2003):
• Springiness of a floor is associated with the sensation of self-gen
erated floor deflection and vibration from a single footstep during
the time of contact between foot and floor surface.
• Vibrational disturbances caused by foot-fall on a floor are charac
terised by perception of floor vibration induced by other persons
than the one that is disturbed.
Figure 7.10 A time history response of timber floor caused by a footfall impact, depicting the forced and
free vibration components; duration of the forced vibration section matches that of the impact duration
7.3.2.4 Damping
Damping is a property which influences vibration amplitudes under
forced vibration and the rate of decay of vibration amplitudes under
free vibration. Damping relates to the dissipation of energy, or to
the conversion of the mechanical energy associated with a vibration
to a form that is unavailable to the vibration. Increased damping
results in rapid decay of a free vibration.
Many floor researchers have attempted to use damping to quantify
the decay of the transient vibration induced by footstep impact, and
hence correlate human response to damping. Despite these attempts,
the problem of how to make use of damping information is largely
unresolved because of the limited knowledge on quantification and
measurements of damping in floor structures having a multiplicity of
vibration modes. Damping values measured by researchers are global
values arising from various sources. Material damping and friction
between components are thought to be the major sources of damping
found in structural systems.
In timber floors, material damping depends on the materials used
for construction. The construction details such as the manner that
components are attached, and the boundary conditions at supports,
contribute to the frictional damping. The material damping is usu
ally small. The major contributor to damping appears to be due to
friction, a factor that has been found in for example cases where
the presence of partitions on floors has provided exceedingly high
damping.
End conditions of bending members can greatly affect damping. It
has been shown that for timber beams with ideally simply supported
ends, the damping ratio associated with the fundamental mode is
about 1 percent, which corresponds approximately to the material
damping. When the same members had ‘clamped’ ends the damping
ratio increased to about 8 percent. For timber joisted floors effective
Karolinska, Solna.
damping ratios are in the range of 1 to 3 percent depending upon
the details of the floor and the mode being considered. Application of
imposed masses (objects) on the surface of a floor can greatly increase
the damping, especially if the system is lightweight or small.
Because of inherent low self-weight, timber floors often do not
exhibit a significant amount of inertial damping. Exceptions can
occur with large systems, especially if they have thick concrete top
ping or support large amounts of imposed mass.
Pt3
a =------ 7.19
48£7
where in this case P = 1 kN. If the most simple approach is used equa
tion 7.17 gives the value of the deflection with the flexural rigidity EI
for one single joist. In most cases however, this will overestimate
the deflection since the joist is only one element in a two-way
load-bearing system. In order to take this into account, the rigidity in
both directions of the floor can be used in order to calculate
the value of the deflection:
P£3
a=K 7.20
48EI
where k is a load distribution factor that can be calculated according
to the following:
Wt(^
P= m 7.22
W
where (EI) is the flexural rigidity of floor in the stiffer direction, that
is along-joist (Nm12/m) and (EI)B is the flexural rigidity of floor in
the direction perpendicular to the stiffer direction, that is across-joist
(Nm2/m), s is the spacing between the joists and l is the span of
the joists, Ohlsson (1984).
Note that other methods exist to perform this calculation and that
various limit values for a are also discussed.
The first criterion is similar to that presented in section 7.3.3.1, page 198,
and Ohlsson (1991) stated that this is a control of the low-frequency
components (< 8 Hz) that are semi-static in nature. The second crite
rion is required to limit the magnitude of the transient response due
to the heel impact of a footstep. As discussed in previous sections,
the impact excites higher frequency components in a timber floor
(8 — 40 Hz) and its response is governed by stiffness, mass and damp
ing of the floor. Believing that contributions to the total motion from
vibration modes with natural frequencies higher than 40 Hz are
insignificant, Ohlsson (1991) proposed the following equation for cal
culating the peak velocity due to unit impulse for a rectangular floor
system simply supported on all four sides, valid for f < 40 Hz:
4(0,4 + 0,6n40)
V = —i--------------- — 7.23
mB£ + 200
where n40 represents the number of eigenmodes with eigenfrequencies
Car park with glulam and CLT frame, Skelleftea.
lower then 40 Hz and is given by:
-1025
7.24
where B is the width of floor (m), l the span (m); m the mass per unit
area (kg/m2), (EI) is the flexural rigidity of floor in the stiffer direction,
that is along-joist (Nm2/m) and (EI)B the flexural rigidity of floor in
the direction perpendicular to the stiffer direction, that is across-joist
(Nm2/m).
Ohlsson (1991) also provides an equation for calculating the funda
mental natural frequency of a floor:
,_2L 7.25
71 2£2 N m
Since the introduction of this method it has been used quite exten
sively and has in many cases shown satisfying results, that is floors
designed according to this method have shown a satisfying
behaviour.
In the second criteria above the damping ratio Z has to be determined
which is a difficult matter. Ohlsson (1991) states that the value on Z
can be 1 percent but also states that higher values can be relevant.
Table 7.1 Vibration classes in office and residential buildings (Talja and Toratti 2006)
Table 7.2 Tentative acceptance limits for vibration classes (Talja and Toratti 2006)
Class a
rms [m/s2]
Peak vertical velocity
3 [mm]
Global displacement of load bearing
member due to 1 kN point load
A < 0,03 < 0,12
floor vibration. For low-frequency floors (f1 < 8 Hz) the recommended
parameter to use is acceleration, because of its good correlation with
subjective ratings.
Talja and Toratti (2006) propose a classification of floors into five
classes, see table 7.1. This classification presumes human walking to
be the designing load case and it is valid for floors in residential and
office buildings. It is based on the sensory perception of a sitting per
son and the human perception of vibrating objects.
This method has the advantage that it gives a tool where the designer
can be more aware of the actual floor performance and on the target
performance to be achieved. Talja and Toratti (2006) also gives tenta
tive limit values on deflections and frequencies for the different
classes, see table 7.2.
Other approaches to modelling and prediction of the dynamic
response can be found in literature but for all of them it can be said
that there are uncertainties and that it is often difficult to find any
one model that can cope with all situations. All the design approaches
are semi-empirical in nature and provide satisfactory solutions for
the particular category of floors on which they are based. None
appears to work entirely satisfactorily when applied to other catego
ries of floors. In many cases it can be said that the best knowledge of
the behaviour of a floor is still obtained by testing it.
Figure 7.11 a) Plywood sub-flooring on timber joists. To the left with wood cross-bracing between joists and to the right with
blocking between joists. b) Plywood sub-flooring with concrete topping on wood I-joists (metal cross-bracing between joists),
Smith (2003).
Timber buildings systems can be used to construct a multitude of 8.1 Structural systems overview 205
buildings. Single-family homes have a relatively simple structure, 8.1.1 Surface elements 205
while greater challenges meet the material when facing multi-storey 8.1.2 Prefabricated Box Units 207
or long-span buildings. Multi-storey timber construction in 8.1.3 Post and beam systems 208
the Nordic countries started around 1995 with the traditional on-site 8.2 Functional requirements and
construction utilizing the 2" x 4" (45 x 95 mm) platform system. system solutions 211
Relatively soon, prefabricated systems emerged. In the Nordic coun 8.2.1 Fire 211
tries, prefabricated systems dominate, while systems with a larger 8.2.2 Acoustics 213
degree of craftsmanship are used in central Europe. Timber building 8.2.3 Horizontal stability 214
systems are common in the Nordic countries and the German speak 8.2.4 Critical moisture level 216
ing countries in central Europe; Germany, Switzerland and Austria
8.3 Detailed design 217
and to some extent Italy. Also, timber building systems exist in
8.3.1 Wall elements 217
the United Kingdom, while timber is less used in countries around 8.3.2 Floor elements 220
the Mediterranean. Globally, USA use timber extensively for housing, 8.3.3 Wall-floor connections 222
the most common technology being a simple 2" x 4" frame stabilised 8.3.4 Services 224
by sheathing. New Zeeland and Australia also apply timber structures 8.3.5 Balconies 226
in housing. LVL is a common framing material both in New Zeeland, 8.3.6 Staircases and elevators 226
Australia and USA. Finally, Japan has a long tradition using timber in
structures. Oftentimes, the great ductility in a timber frame ensures
good racking resistance under earthquake loading.
• Post and beam systems: are sometimes named big frame structures
and are very common for industrial and commercial buildings.
These systems are also used for buildings or parts of buildings,
where the load-carrying structure is visible. The post and beam sys
tem is a system based on a structural grid of beams and columns,
typically with pinned connections. The structure is instead stabi
lised through diagonal bracing or shear walls. Post and beam sys
tems are suitable for long-span structures and are used for arenas.
Figure 8.3 a) Structural system with cross-laminated elements. b) X-Lam, Massivtrahandboken (2006).
the walls in modular systems since the outer walls of the module are
used as vertical load-carrying components.
Utilising this kind of system gives the opportunity to control
the whole building process from architectural and structural design
to completion. The architectural design of the buildings built with
modules needs to be carried out with knowledge of the building sys
tem and the limitations and possibilities in the production system.
Modular systems are suitable for student housing, hotels and housing
for the elderly, but is also adaptable to condominiums or tenancies.
Each module can comprise one or a few rooms, or even the whole
apartment in the case of student apartments. The size of the modules
is determined by manufacturing, handling and transportation
limitations.
Figure 8.9 a) The post and beam system "Tra8", Askims torg.
b) Hockey hall, Ekvallen.
c) Bandy hall, Nassjo.
Such long-span structures are not used for housing, which means
that most of the functional requirements in place for housing are not
viable for this type of structures. Acoustics is seldom a problem;
moisture safety and thermal insulation are currently poorly
addressed. Load-carrying capacity and fire are the most prominent
requirements to address for the structure.
To reach an economical structure, optimisation of the load-carry
ing capacity is therefore a must, which includes efficient design of
10°
a = 3 - 10°
Double pitched beam with curved underside Three-hinged portal frame with curved haunches
a >14
Three-hinged tied trussed rafter Three-hinged portal frame with knee bracing
8.2.1 Fire
The European codes for fire safety in timber buildings have been
revised and handbooks published, Fire Safety in Timber Buildings
(2010). Buildings should limit the risk related to fire, limit the spread
of fire and fire gases inside the building, and allow the inhabitants to
escape from the building or to be evacuated or saved. The safety of
firemen is also taken into consideration.
The building fire documentation that is mandatory for all build
ings should include: fire safety classes of used materials and ele
ments, division into fire cells, evacuation plan, ventilation during
fire, and a description of used technical solutions, including any
additional installation like sprinklers or fire detection systems. It is
recommended to integrate fire safety design in the very early stages
of building design.
Structural elements, materials, claddings and surface finishes,
depending on their function are assigned one of the following prop
erties, EN 1995-1-2:
• R — load-bearing capacity
• RE — load-bearing capacity and integrity
• REI — load-bearing capacity, integrity and insulation
• E — integrity
• EI — integrity and insulation.
8.2.1.1 Timber
The wood material itself burns and is destroyed in a fire. However, it
burns in a controlled manner and quite slowly. A protective layer of
coal is created on the surface when a wooden piece is subjected to
fire. If the wooden piece is thick, the core of the wood will remain
intact for a good deal of time. The charring rate is of the order
0,7 mm/min. In the case of large cross sections (for example glulam),
the slow charring rate makes it possible to evacuate the building
before collapse. The glue used for glulam can sustain fire well.
Design for fire when looking at large cross sections (as those used in
long-span structures) is made by first calculating a design load using
the load combination for fire in EN 1990. Then, the reduction of
the cross section is calculated corresponding to the required fire
resistance time (for example R 60), using the design charring rate
specified in EN 1995-1-2. If the structure can sustain the load with
Ulls hus, Uppsala.
the remaining cross section, the fire resistance is sufficient.
When considering light-weight structures, the cross section of
wood is much smaller (typically 45 mm on the short side). This
dimension is too small to sustain a fire during for example 60 min
utes and therefore the structure needs to be completed with some
type of cladding to increase the fire resistance.
8.2.1.2 Sheathing
In light frame timber walls, the fire resistance depends in decreasing
importance upon the following parameters, Fire Safety in Timber
Buildings (2010):
• the sheathing firstly exposed to fire
• the sheathing secondly exposed to fire
• cavity insulation protecting the sides of the timber member
• type of insulation in cavity, stone wool is generally better than
glass wool.
8.2.1.3 Sprinklers
Sprinklers are a type of active fire protection. They are used to a vary
ing extent in different countries; some prescribe them in buildings
higher than for example 4 storeys while some do not. There are basi
cally two types of sprinkler systems: industrial and residential.
Residential sprinklers are an important novelty since most fires take
place in residential buildings. The residential sprinklers are quite
easy to assemble in the building and using them gives a basis for risk
assessment that may lead to a technical exchange in fire safety
design, see Ostman et al. (2002). Some examples of successful applica
tions of technical exchange when using sprinkler systems are:
8.2.2 Acoustics
Acoustic regulations for housing, educational buildings and some
other building categories now exist in most countries in Europe, but
findings from comparative studies show that extent and strictness as
well as descriptors vary considerably across Europe. The acoustic per
formance areas dealt with are e.g. airborne and impact sound insula
tion, reverberation time, traffic noise, service equipment noise.
Comparing countries, there is in general no consistency of contents,
structure or enforcement of acoustic regulations, see Rasmussen
(2018).
Timber structures are treated as light due to their relatively low
mass in comparison to concrete and steel structures. An overview of
the state of the art regarding acoustics in timber frame buildings is
Figure 8.13 Roof used as weather protection during erection of the structure
a)
25 Wood cladding
34 Ventilation space between vertical battens
Wind protection
195 Thermal insulation between vertical studs s600
83 CLT
Vapour barrier
45 Thermal insulation between horizontal studs s600
15 Gypsum board (one or two)
15 Gypsum board
94 CLT
2 x 30 Sound insulation board
94 CLT
15 Gypsum board
b)
The example in figure 8.16 b) with double CLT panels with sound
insulation boards at the centre provides good sound reduction for air
borne sound (61 dB).
Apartment separating walls in timber tend to become quite thick,
which reduces the area possible to sublet for the client. Thick walls
can decrease the return on investments, what also arises for modular
buildings, where the wall structure is double in the building system.
Outer corners in walls also needs to be connected, see figures 8.17
and 8.18:
External walls are thicker than internal walls to provide good insu
lation. The thickness of external walls increases with increasing
demands on energy efficiency. External walls are subject to several
functional requirements; thermal insulation, moisture barrier, sound
insulation from sounds in the outer environment, horizontal, and
vertical load-carrying capacity. External walls of solid wood panels
need to be completed with insulation material, thus implying that
a regular light-weight wall must be built outside the actual load-car
rying structure, see figure 8.17. Today there are also systems where
the facade is hung up in the primary structure without extra studs,
what also reduces thermal bridges. Light frame walls are on the other Figure 8.18 Horizontal section of external wall
hand not quite as strong as solid wood walls. The connection between connection in a light frame system. The wall cross section is
light frame walls in the outer corner can be realised as in figure 8.18. the same as in figure 8.15 a), completed with corner details.
22 Particle board
45 Wood stud (45 x 120)
12 Particle board
245 Wood joist +
120 Insulation
145 Wood joist
28 Wood batten
15 Fire resistant gypsum board
a) b)
Figure 8.21 Illustration of a) cross-laminated element, b) cassette element and c) composite floor design.
lining 28 x 70 si 200 mounted on site is cut after mounting the suspended ceiling to the walls
Figure 8.22 Example of a floor cassette that can be used for span lengths up to 6,5 m
The deck is made of a 3-layered cross laminated plate, supported by glulam beams with bottom flanges with a suspended ceiling.
result in a weaker joint with a risk for increased deflection (note that
the connectors are not visible in figure 8.24 b). It can also hang only on
the flanges, which often is the case for cassette floors, see figure 8.22,
page 221.
In many cases the structural details are not commonly available.
Solutions can be found only as examples of existing structures. In
the wall-floor connection, many functional requirements must be
met. Firstly, the vertical load bearing must be granted. In wood,
the compression stresses perpendicular to the grain must be moni
tored for the vertical load. Secondly, the horizontal resistance must
be sufficient; any connections must be able to transfer horizontal
forces between stories. If the floor is placed on top of the wall as in
figure 8.24 a), the horizontal load must pass the floor element.
1. Horizontal stud
1 2. Wind protection
3. Sole plate
4. Edge beam
6 5. Top plate
2 6. Thermal insulation between vertical studs
7
7. Gypsum board
8 8. Vapour barrier
a)
1. Wind protection
2. Sole plate
3. Top plate
4. Thermal insulation between vertical studs
5. Vapour barrier
1
6. Gypsum boards
7. Edge beam
2
J
3
4
1
b)
Figure 8.24 a) Example light-frame structure with floor element placed on top of the walls.
b) Example light-frame structure with floor element hanging from the walls.
8.3.4 Services
Timber structures are definitely advantageous compared to concrete
and steel structures regarding the services, primarily due to easiness
of making openings on site. On the other hand it is always more
complicated to integrate the services in prefabricated building com
ponents than during on-site construction. Clients always demand
a documentation of the installation system when a building system is
used. Generally, in prefabricated building systems, it is favourable to
Figure 8.26 Hanger, MFB Handbok (2010) lead the services vertically to reduce the number of passages between
fire cells, see Lennartsson (2009). In multi-storey residential buildings
------- 4
------- 5
1. Multilayer cross-laminated timber
2. 145 Thermal insulation
3. Wind protection Element joint
4. 34 x 45 Vertical battens
producing a ventilation space 6
5. Wood cladding
6. Steel support reducing flanking
sound transmission
7. 2 x 13 Gypsum board
| | Dwelling
| | Shaft
| | Service central
• Canalisation with
connection point room
Figure 8.28 a) Single and b) multiple shafts for services, Lennartsson (2009).
a)
For a system based on panel elements (for example CLT), the service
layer can be placed either on top in a floating floor or under the solid
wood panel in the suspended ceiling. The choice of solution is strictly
connected to the step sound insulation. If the ceiling is not included
in the floor element, the services can be placed in this space before
finishing, figure 8.30. If the ceiling is integrated in the floor element,
services are installed at the factory. With cassette floor elements,
the height of the floor does not need to be increased because of ser
vices since the spaces between joists can be used for this purpose,
figure 8.31.
8.3.5 Balconies
a) An example of a solution of a 1 200 mm x 1 200 mm balcony for
a multi-storey timber building is presented in figure 8.32, page 227.
The solution consists of a CLT plate placed in a steel frame anchored
in the CLT structure, see Martinsons. However, many different system
solutions for timber balconies have been developed the last years, see
for example Pousette et al. (2011).
1. CLT
2. Sealing layer
3. Pressure-treated battens
4. Pressure-treated decking
5. Z-profile of two welded steel L-profiles
6. Drip plate
7. Steel L-profile
8. Facade ribs
9. Lintel
10. Fixing with going-through rod
11. Drip plate
12. Tension steel tie
b-b
Structural systems
for infrastructure
Anna Pousette
There are many types of timber bridges, see table 9.1. The bridge deck
typically consists of beams or a deck plate. For longer spans the bridge
deck can be complemented by other structural parts. Different bridge
types such as trusses, arches, cable stayed bridges and suspension
bridges can be used for both road bridges and pedestrian bridges, see
TraGuiden (www.traguiden.se).
Beams 5 - 30 m
& &
£
Kings post 10 - 50 m
__ £
&
Strut frame 20 - 40 m
— Truss 10 - 40 m
oo
>n---------
Arch 20 - 70 m
—
Cable-stayed 20 - 130 m
( J &
Suspension 20 - 130 m
77 77 *7
Kings post truss bridges: have vertical rods that transfer loads from
the bridge deck to diagonal members at the sides of the bridge deck.
Kings post bridges can have large spans with a relatively small height
of the bridge deck.
Strut frames: are old building systems that have diagonal compres
sion struts under the bridge deck. The construction below the deck
enhances the load-bearing capacity of the bridge deck or beam and
transfers the compressive forces to the abutments or walls.
Trussed beams and trusses: can span longer with less material than
homogeneous beam sections, but require larger structural depth.
These bridges are built of glulam and are sometimes combined with
tie rods of steel. They have many joints, which are very important for
the strength, stiffness and durability of the structure. The trusses can
be placed under the bridge deck or at the sides of the bridge deck.
When possible, trusses below the bridge deck are favourable because
several trusses can then be used, the secondary system will be sim
pler and the trusses will be protected against weather. When trusses
are placed at the sides of a road bridge the depth of the truss should
be big enough for a horizontal wind bracing system to be installed
above the traffic lanes. The truss chords can also be designed with
curved elements.
Trusses at the sides of the deck are used for pedestrian bridges with
a width of 2 — 3 m and trusses with depth of about 1 — 1,5 m can act
as bridge railing together with complementary parts. Experience
shows that protection against moisture is important, but usually for
aesthetic reasons the cladding does not cover the entire trusses. Yet
for truss bridges, all exposed upper surfaces of the lower chord and
the lower diagonal joints should be covered using metal plate
cladding.
Arches: are usually made of glulam. The lamellas are bent during
the gluing process and they can be performed with different radii.
Due to limitations and efficiency of production and transportation
the glulam arches are often made as 3-hinge arches. The arches can
be placed below or above the bridge deck. Arches above the deck are
placed at the sides and are combined with crossbeams under the deck
and steel hangers. The joints are critical to the load-bearing capacity
and durability, and the arches should be protected with cover plates
and cladding. The arches can be treated with chemical impregnation
or be untreated if they are well covered.
Railing
Protective layer
and asphalt pavement
Cover plate
Cladding
Timber deck plate
where:
bw,middle should be calculated according to figure 9.8.
a is given in the standard, for example 0,3 m for
a stress-laminated deck plate. This beam is loaded
with distributed load and wheel loads and the required
deck thickness can be calculated.
When wheel loads are placed on the deck, the transverse bending
produces a tendency of gaps forming between the laminations on
the lower side of the deck and the transverse shear creates a tendency
of vertical slip of the laminations, see figure 9.9. A sufficient level of
compressive pre-stress between the laminations prevents both actions.
For transverse bending, the compressive stress directly counteracts
the tension on the lower side of the deck. Vertical slip is prevented by
Transverse bending
the friction between the laminations resulting from the compressive
pre-stress. Maintaining an adequate level of pre-stress is the most
■
important issue regarding a stress-laminated deck plate. If
the pre-stressing forces are reduced over time due to shrinkage or
creep, it is always possible to re-stress them. The loss of pre-stressing
force is considerable during the first period after construction, but
after re-stressing a number of times during the service life the force
seems to stabilize as the moisture content also stabilizes. However,
Transverse shear (rolling shear) the pre-stressing force will always vary significantly with temperature.
The long-term pre-stressing forces shall be high enough to ensure
Figure 9.9 Actions of wheel loads that no inter-laminar slip occurs. The maximum allowed shear force
on stress-laminated decks
caused by vertical and horizontal actions depends on the coefficient
of friction, the minimum long-term residual compressive stress due
to prestressing, and the thickness of the plate. The coefficient of fric
tion is determined based on wood species, roughness of contact sur
face, treatment of the timber, moisture content, and residual stress
level between laminations. In areas subjected to concentrated loads,
the minimum long-term residual compressive stress due to pre-stress
ing between laminations should not be below 0,35 N/mm2.
According to Eurocode EN 1995-2 the long-term residual pre-stress
ing can normally be assumed to be greater than 0,35 N/mm2, provided
that the initial pre-stress is at least 1,0 N/mm2, the moisture content
of the laminations at the time of pre-stressing is not more than 16 %,
and the variation of in-service moisture content in the deck plate is
limited by an adequate protection layer.
The resulting pre-stressing forces should act centrally on the tim
ber cross-section. The compressive stress perpendicular to the grain
during pre-stressing in the contact area of the anchorage plate shall
be verified. As the pre-stressing forces at the edge can be too high for
the softwood of the deck, hardwood parts are often used for distribu
tion of the pre-stressing.
Stress-laminated bridges are common for both pedestrian bridges
and road bridges. There are three kinds of stress-laminated structures:
deck plates, T-beam bridges and box-beam bridges. Deck plates are
built as either simply supported or continuous over several supports.
T- and box-beam bridges are usually simply supported with joints at
each support. T- and box-beam bridges are used for longer spans. This
design becomes more complex, see figures 9.10 and 9.11. Transfer of
shear forces between web and flanges as well as deformations for
asymmetrical loads are significant in the design of T-beams and
box-beams.
T-beam bridges and box-beam bridges become somewhat deeper
than a deck plate with the same span, but they are built with less
material. Suitable spans and an approximate indication of required
depths are shown in table 9.2, page 235, for simply supported bridges
with a deck width of about 3 m. The depth depends on the load,
the paving, the wood material, and the width. For T- and box-beam
bridges it also depends on the number of webs and their width and
material.
There are also other types of decks manufactured of planks or glu
Figure 9.11 Stress-laminated box-beam bridge lam beams placed side by side and nailed or glued together, or of
Fastener
cross-laminated timber (X-Lam) or laminated veneer lumber (LVL).
Concrete
X-Lam consists of at least three layers of crosswise glued boards (most
commonly of softwood), but also five or seven layers are used, see Non-structural
also section 2.7, page 47. The X-Lam boards can be combined with intermediate layer
glued laminated beams to form a T-section. The connection between
Wood
the bridge deck and the beams is important to transmit the shear
forces.
The idea of timber-concrete composites is to combine the high ten
sile strength of timber with the high compression resistance of con Figure 9.12 Timber-concrete composite deck, with
crete. The first applications were in structural systems for new build intermediate layer
ings and in renovation. The shear forces can be transmitted by
mechanical connectors or by direct contact between timber and con
crete (grooved connection). The stress distribution depends on
the rigidity of the connectors.
Timber-concrete connections in composite beams can be designed
as laterally loaded dowel-type fasteners. Where there is an intermedi
ate non-structural layer between the timber and the concrete (for
example for formwork), see figure 9.12, the strength and stiffness
parameters should be determined by applying a special analysis or by
tests.
For grooved connections, the shear force should be taken through
direct contact pressure between the wood and the concrete cast in
the groove. Sufficient resistance of the concrete and timber parts of
the connection should be verified. The concrete and timber parts must
be held together to avoid separation and the connection should be
designed for a small tensile force between the timber and the concrete.
Figure 9.17 Wind load on noise barrier (highest load in zone A near the end)
Symbols
Symbols used in EN 1995-1-1.
Fax,Rd Design value of axial withdrawal capacity of Lnet,t Net width of the cross-section perpendicular
the fastener to the grain
Rv,drl Design racking capacity of a wall diaphragm ft Design tensile strength of flange
Vu, Vl Shear forces in upper and lower part of beam fhead,k Characteristic pull-through parameter for nails
with a hole
f, Fundamental frequency
wy Section modulus about axis y
fm,k, Characteristic bending strength
Xd Design value of a strength property
fm,y,d„ Design bending strength about the principal
Xk Characteristic value of a strength property y-axis
Latin lower case letters fm,z,d„ Design bending strength about the principal
z-axis
a Distance
fm,a,d„ Design bending strength at an angle a to
a Spacing, parallel to grain, of fasteners within
the grain
one row
f:.0.d Design tensile strength along the grain
Minimum end distance to the centre of gravity
a1,CG
of the screw in each timber member f:.0.k Characteristic tensile strength along the grain
ad Spacing, perpendicular to grain, between rows ft,90,d Design tensile strength perpendicular to
of fasteners the grain
a2,CG Minimum edge distance to the centre of f:.w.d Design tensile strength of the web
gravity of the screw in each timber member
f• Characteristic tensile strength of bolts
a3,c Distance between fastener and unloaded end
fv,0,d Design panel shear strength
a3,t Distance between fastener and loaded end
fv,ax,a,k, Characteristic withdrawal strength at an angle
a4,c Distance between fastener and unloaded edge a to grain
a4,t Distance between fastener and loaded edge fv,ax,90,k Characteristic withdrawal strength
perpendicular to grain
abow Maximum bow of truss member
fv,d„ Design shear strength
abow,perm Maximum permitted bow of truss member
h Depth; height of wall
adev Maximum deviation of truss
hap Depth of the apex zone
adev,perm Maximum permitted deviation of truss
hd Hole depth
b Width
he Embedment depth; loaded edge distance
bi Width of panel i or wall /
hefr Effective depth
bnet Clear distance between studs
hf,c Depth of compression flange
bw Web width
hf,t Depth of tension flange
d Diameter; outer thread diameter
hrl Distance from lower edge of hole to bottom of
d, Diameter of centre hole of connector; inner
member
thread diameter
hru Distance from upper edge of hole to top of
dc Connector diameter
member
def Effective diameter
hw Web depth
dh Head diameter of connector
i Notch inclination
fh,i,k Characteristic embedment strength of timber
kc,y, kc,z Instability factor
member i
kcr Crack factor for shear resistance
fa,0,0 Characteristic anchorage capacity per unit area
for a = 0° och p = 0° kcrit. Factor used for lateral buckling
fa,90,90 Characteristic anchorage capacity per unit area kd Dimension factor for panel
for a = 90° och p = 90°
kdef Deformation factor
fa.^ Characteristic anchorage strength
k.dis Factor taking into account the distribution of
fax,kt Characteristic withdrawal parameter for nails stresses in an apex zone
f,0.d Design compressive strength along the grain kf,1, kf,2, kf,3 Modification factors for bracing resistance
f,f,c,d„ Design compressive strength of flange ki,q Uniformly distributed load factor
kv Reduction factor for notched beams p Angle between the grain direction and the
force for a punched metal plate
kvol> Volume factor
Pc Straightness factor
ky eller kz Instability factor
Y Angle between the x-direction and the timber
l a,min Minimum anchorage length for a glued-in rod
connection line for a punched metal plate
l Span; contact length Partial factor for material properties, also
YM
Distance from a hole to the centre of accounting for model uncertainties and
lA
the member support dimensional variations
Distance from a hole to the end of the member Az Slenderness ratio corresponding to bending
lv
about the z-axis
lZ Spacing between holes
,
Arel,y Relative slenderness ratio corresponding to
m Mass per unit area bending about the y-axis
Bibliography
Chapter 1 - Introduction to design and design process
Engstrom, B.: Conceptual Design — Methodology. Course literature on structural analysis and design.
Department of Structural Engineering. Chalmers University of Technology, Gothenburg, Sweden, 2002.
Johansson, G., Perstorper, M., Kliger, R.: Kvalitetskrav pa byggnadsvirke. Kvalitetskrav pa byggnadsvirke.
(In Swedish). BFR R105:1990, Stockholm, 1990.
Kroll, E., Condoor, S., Jansson, D.: Innovative Conceptual Design — Theory and Application of Parameter
Analysis. Cambridge University Press. Cambridge, UK, 2001. ISBN 0521778484.
Niemeyer, S.: Conceptual Design in Building Industry. Master’s Thesis 03:6. Department of Structural
Engineering. Chalmers University of Technology, Gothenburg, 2003.
Paik, J.K., Melchers, R.E.: Condition assessment of aged structures. Woodhead Publishing Limited, 2008.
ISBN 1 84569 334 5.
Thelandersson, S., Isaksson, T., Fruwald, E., Toratti, T.,Viitanen, H., Grull, G., Jermer, J., Suttie, E.:
Bestandighet for utomhustra ovan mark. Guide for utformning och materialval. Report TVBK-3060.
Faculty of Engineering, Lund University, 2012.
Anon (2005a): European Forest Sector Outlook Study — 1960 — 2000 — 2020 Main report. United Nations
publications, 2005. ISSN 1020 2269.
Anon (2005b): Global Forest Resources Assessment 2005. FAO Forestry Paper 147. Food and Agriculture
Organization of the United Nations, Rome, 2005.
Bodig, J., Jayne, B.: Mechanics of wood and wood composites. Van Nostrand Reinhold company Inc.
New York, 1982. ISBN 0-442-00822-8.
Dackling, M.: Den svenska travarumarknaden. Forbrukning inom delsektorer 1995 — 2000. (The Swedish
market for sawn timber. Use within different industry segments, in Swedish). Institutionen for industriella
produktionssystem, avdelningen for Skog- och trateknik, Vaxjo universitet, Vaxjo, 2002.
Dinwoodie, J.M.: Timber: Its nature and behaviour. Second edition. E & FN Spon, London, 2000.
ISBN 0-419-23580-9.
Hankinson, R.L.: Investigation of crushing strength of spruce at varying angles of grain,
Air Force Information Circular No. 259, U. S. Air Service, 1921.
Hearmon, R.F.S. Paton J. M.: Moisture content changes and creep of wood.
Forest products journal 14 (8) pp. 357 — 359, 1964.
Hill, C.A.S.: Wood Modification: Chemical, Thermal and Other Processes. John Wiley&Sons,
Chichester, England, 239 pp, 2006. ISBN 0-470-02172-1.
Hoffmeyer, P.: Wood as a building material (A4), in Timber Engineering STEP 1, Blass, H.J.,
Aune P., Chooet B.S., Editors. Almere, Centrum Hout, the Netherlands, 1995.
Hoffmeyer, P.: Strength under long-term loading, in Timber Engineering, S. Thelandersson and H.J. Larsen,
Editors. Wiley: Chichester. ss.131-152, 2003. ISBN 0-470-84469-8.
Hon, D.N.S., Shiraishi, N. (ed).: Chemical modification of wood, Handbook on Wood and Cellulosic Materials,
Marcel Dekker, Inc., New York, NY, Chapter 15, 703 — 756, 1991.
Johansson, C.-J.: Grading of Timber with Respect to Mechanical Properties, in Timber Engineering,
S. Thelandersson, Larsen, H. J., Editors. Wiley: Chichester. ss. 23 — 45, 2003. ISBN 0-470-84469-8.
Skogsindustrierna, The Swedish Forest Industries: Facts and Figures.
Timell, T.E.: Compression wood in gymnosperms. New York; Springer-Vlg, Cop, 1986. ISBN 0-387-15715-8.
Blass, H.J.: Columns. In “Timber Engineering STEP 1”. Ed. Blass et. al., Centrum Hout, the Netherlands, 1995.
Carling, O.: Dimensionering av trakonstruktioner (Design of Timber Structures), in Swedish,
AB Svensk Byggtjanst & Tratek, Stockholm, Sweden, 1992.
Danielsson, H.: The strength of glulam beams with holes. A Probabilistic Fracture Mechanics Method and
Experimental Tests. Division of Structural Mechanics. Report ISRN LUTVDG/TVSM--09/3069--SE (1 — 124), 2008.
Gustafsson, P.J.: Notched beams and holes in glulam beams. In “Timber Engineering STEP 1”.
Ed. Blass et. al., Centrum Hout, the Netherlands, 1995.
Martensson, A.: Short and long term deformations of timber structures. In “Timber Engineering”.
Ed. Thelandersson S., Larsen H.J., Wiley & Sons, Chichester, 2003.
Piazza, M., Tomasi, R., Modena, R.: Strutture in Legno (Timber Structures), in Italian,
Hoepli, Milan, Italy, 2005.
Porteous, J., Kermani, J.: Structural Timber Design to Eurocode 5, Blackwell Science Ltd, Oxford, UK, 2007.
Stalnaker, J.J., Harris, E.C.: Structural Design in Wood. Van Nostrand Reinhold, New York, 1989.
Timoshenko, S., Gere, J.M.: Theory of Elastic Stability (2nd edition). McGraw Hill International Book Company,
Singapore, 1963.
Thelandersson, S., Larsen, Timber Engineering. John Wiley & Sons Ltd. West Sussex, UK. 3.41, 2003.
Blass, H.-J.: Load distribution in nailed joints. In proceedings from the International Council for
Research Innovation in Building and Construction, Working Commission CIB-W18, paper 23-7-2,
Lisbon, Portugal, September 1990.
Blass, H-J.: Joints with Dowel-Type Fasteners. Timber Engineering. Larsen H.J., Thelandersson, S. Eds.,
Wiley & Sons, 2003. ISBN 0-470-84469-8.
Cramer, C.O.: Load Distribution in Multiple-Bolt Tension Joints. Journal of the Structural Division
94(ST5):1101-1117, 1968.
Gustafsson, P.-J., Serrano, E.: Glued-In Rods for Timbers Structures (GIROD): Final Report for WP1 —
Development of a calculation model, Lund University, 2001.
Johnsson, H.: Plug Shear Failure in Timber Connections — Avoiding Brittle and Promoting Ductile Failures.
Doctoral thesis 2004:03, Lulea University of Technology, 2004.
Kuipers, J., Van Der Put, T.A.C.M.: Betrachtungen zum Bruchmechanismus von Nagelverbindugen.
In Ingenieurholzbau in Forschung und Praxis, Eds Ehlbeck J. and Steck G., Bruderverlag Karslruhe, 1982.
ISBN 3-87104-049-5.
Larsen, H.J. och Riberholt, H.: Traekonstruktioner (in Danish). SBI-anvisning 194,
Statens Byggeforskningsinstitut, 1999.
Nielsen, J.: Trusses and Joints with Punched Metal Fasteners. Timber Engineering.
Larsen H.J., Thelandersson S. Red., Wiley & Sons, 2003. ISBN 0-470-84469-8.
Tlustochowicz, G., Serrano, E., Steiger, R.: State-of-the-art review on timber connections
with glued-in steel rods. Materials and Structures 44(5):997-1020, 2011.
Foschi: Rolling shear failure of plywood in structural components. Forest Products Laboratory, Madison.
Information report VP-x-67, 1970.
Granholm, H.: Om sammansatta balkar och pelare med sarskild hansyn till spikade trakonstruktioner.
(On composite beams and columns with particular regard to nailed timber structures). Transactions,
Chalmers University of Technology, Rep. No. 88, 1949. (In Swedish)
Lekhnitsky: Anisotropic plates. Gordon & Breach Science Publishers, New York, 1968.
Newmark, N.M., Siess, C.P, Viest, I.M.: Tests and analysis of composite beams with incomplete interaction.
Proceedings, Society for Experimental Stress Analysis 9(1): 75-92, 1951.
Oduyemi, T.O.S., Wright, H.D.: Partial interaction analysis of double skin composite beams.
Journal of Constructional Steel Research 19(4): 253-283, 1991.
StBK-N5: Tunnplatsnorm. AB Svensk Byggtjanst, Stockholm, 1980. (In Swedish)
Stussi, F.: Zusammengesetzte vollvandtrager. (Built-up girders). International Association for Bridge
and Structural Engineering, IABSE, Vol. 8, pp. 249-269. (In German)
Thielgard, Larsen: Limede I-bj^lker af tra. Aalborg University, inst. for Bygningsteknik.
Report 7804, 1978. (In Danish)
Belluzzi, O.: Scienza delle Costruzioni, vol. IV (Structural Mechanics, vol IV), in Italian,
Zanichelli, Bologna, Italy, 1966.
Carling, O.: Dimensionering av trakonstruktioner (Design of Timber Structures), in Swedish,
AB Svensk Byggtjanst & Tratek, Stockholm, Sweden, 1992.
Caironi, M., Bonera, L.: Il legno lamellare, il calcolo (Glued laminated timber, design), in Italian,
Habitat Legno editor, Edolo (BS), Italy, 1989.
Gross, H., Crocetti, R.: Limtrahandbok Del 1 — 3, in Swedish, Svenskt Tra, Stockholm, 2016.
Holzbau Handbuch: Reihe 2 Teile 2 12 folge 1, Verbande und Abstutzungen — Grundlagen, Regelnachweise
(Connections and bracing — basics, design guide), in German, Informationdienst Holz, Deutschen
Gesellschaft fur Holzforschung e. V., Munich, Germany, 2000.
Piazza, M., Tomasi, R., Modena, R.: Strutture in Legno (Timber Structures), in Italian,
Hoepli, Milan, Italy, 2005.
Porteous, J., Kermani, J.: Structural Timber Design to Eurocode 5, Blackwell Science Ltd, Oxford, UK, 2007.
Schodek, D.L.: Structures, 2nd edition, Prentice-Hall, New Jersey, 1992.
Schueller, W.: Horizontal-Span Building Structures, John Wiley & Sons, New York, 1983.
Stalnaker, J.J., Harris, E.C.: Structural Design in Wood. Van Nostrand Reinhold, New York, 1989.
Sorensen, J.D., Dietsch, P., Kirkegaard, P.H., Kohler, J.: Design for Robustness of Timber Structures.
COST Action E55 “Modelling of the Performance of Timber Structures”. European Science Foundation,
Brussels, 2011. ISBN 978-3-8322-9949-1.
Timoshenko, S., Gere, J.M.: Theory of Elastic Stability (2nd edition).
McGraw Hill International Book Company, Singapore, 1963.
Thelandersson, S., Larsen, H.J.: Timber Engineering. John Wiley & Sons Ltd. West Sussex, UK, 2003.
Vessby, J.: Analysis of shear walls for multi-storey timber buildings. PhD thesis, Linnaeus University,
Vaxjo, Sweden, 2011.
Winter, G.: Lateral bracing of columns and beams. Proc. ASCE 1958,84 (ST2), 1561-1-1561-22, 1958.
Yura, J.: Winter’s bracing approach revisited. Engineering Structures, Vol. 18, No. 10, pp. 821 — 825,
Elsevier Science Ltd, UK, 1996.
Hu, L.J., Chui, Y. H., Onysko, D.,: Vibration serviceability of timber floors in residential construction,
Prog. Struct. Enging. Mater, 3:228 — 237, 2001.
Martensson, A.: Short and Long-term Deformations of Timber Structures, Timber Engineering,
Eds: Larsen H.J., Thelandersson S., Wiley & Sons, 2003.
Ozelton, E.C., Baird, J.A.: Timber Designers’ Manual; Blackwell Science Ltd, 2006.
Porteous J., Kermani, A.: Structural Timber Design, Blackwell Science Ltd, 2007.
Ranta-Maunus, A., Kortesmaa, M.: Creep of timber during eight years in natural environments.
WCTE2000 Conference. Whistler, CA, 6 p, 2000.
Smith, I.: Vibrations of Timber Floors: Serviceability Aspects. Timber Engineering,
Eds: Thelandersson, S., Larsen, H.J., John Wiley & Sons Ltd. West Sussex, UK, 2003.
Forssen, J., Kropp, W., Brunskog, J., Ljunggren, S., Bard, D., Sandberg, G., Ljunggren, F., Agren, A., Hallstrom,
O., Dybro, H., Saint-Gobain, I., Larsson, K., Tillberg, K. Jarnero, K., Sjokvist, L-G., Ostman, B., Hagberg, K.,
Blomkvist, A., Olsson, A., Ekstrand C-G., Johansson, M.: Acoustics in wooden buildings. State of the art 2008.
Vinnova project 2007-01653. SP Rapport 2008:16, SP Hallbar Samhallsbyggnad, Stockholm, 2008.
Gross, H., Crocetti, R.: Limtrahandbok Del 1 — 3, in Swedish, Svenskt Tra, Stockholm, 2016.
KLH: www.klh.at
Lennartsson, M.: Modularity in Industrialised Timber Housing. A Lean Approach to Develop
Building Service Systems. Lulea University of Technology, Licentiate thesis, 2009.
Lindbacks: www.lindbacks.se
Ljunggren, S.: Ljudisolering i trahus — en handbok for konstruktorer, in Swedish.
SP Swedish Technical Research Institute, Stockholm, 2011.
Natterer, J., Hamm, J., Favre, P.: Composite wood-concrete floors for multi-storey buildings.
Proc. Of the International Wood Engineering Conference, 431 — 435, 1996.
Nordlund, P.: Horisontalstabilisering av hoga trahus via trapphuset (in Swedish).
Master thesis, Lulea University of Technology, 2010.
The Open Timber Construction System: The Open Timber Construction System
— Prefabricated Element Structures. WoodFocus, Finland, 2003.
TraGuiden, The Swedish Forestry Industries Federation (Skogsindustrierna), www.traguiden.se, (In Swedish).
Tyrens & Derome: Traplattformen — en plattform for industriellt trabyggande. CBBT, 2011.
Vessby, J.: Shear walls for multi-storey timber buildings. Licentiate thesis, School of Technology and Design,
Vaxjo University, Sweden, 2008.
Ostman, B., Arvidson, M., Nystedt, F.: Boendesprinkler raddar liv — Erfarenheter och
brandskyddsprojektering med nya mojligheter (in Swedish), SP Rapport, SP Hallbar Samhallsbyggnad, 2002.
Pousette, A., Jacobsson, P., Gustafsson, M., Horttanainen, J., Dahl, K.:
Stress Laminated Bridge Decks, Part II. Tratek — Swedish Institute for Wood Technology Research,
Nordic Timber Bridge Project, Nordic Wood, 2002.
Pousette, A., Fjellstrom, P.-A.: Inspection of timber bridges, Proceedings of 7th
International conference on short and medium span bridges, Montreal, Canada, 2006.
TraGuiden, The Swedish Forestry Industries Federation (Skogsindustrierna), www.traguiden.se, (In Swedish).
EN 1075:2014 Timber structures — Test methods — Joints made with punched metal plate fasteners.
EN 1382:2000 Timber structures — Test methods — Withdrawal capacity of timber fasteners.
EN 1383:2000 Timber structures — Test methods — Pull through resistance of timber fasteners.
EN 1990:2002/A1:2005/AC:2009/AC:2010 Eurocode 0: Basis of Structural Design.
EN 1991-1-2:2002/AC:2009/AC:2013/AC2:2013 Eurocode 1:
Actions on structures — Part 1 — 2: General actions — Actions on structures exposed to fire.
EN 1991-1-3:2003/AC:2009 Eurocode 1:
Actions on structures — Part 1 — 3: General actions — Snow loads.
EN 1991-1-4:2005/AC:2009/AC:2010/A1:2010 Eurocode 1:
Actions on structures — Part 1 — 4: General actions — Wind actions.
EN 1991-1-5:2003/AC:2009 Eurocode 1:
Actions on structures — Part 1 — 5: General actions — Thermal actions.
EN 1991-1-6:2005/AC:2008/AC:2013 Eurocode 1:
Actions on structures — Part 1 — 6: General actions — Actions during execution.
EN 1991-1-7:2006/AC:2010/A1:2014 Eurocode 1:
Actions on structures — Part 1 — 7: General actions — Accidental actions.
EN 1991-2:2003/AC:2010 Eurocode 1:
Actions on structures — Part 2: Traffic loads on bridges.
EN 1993-1-5:2006/AC:2009 Eurocode 3:
Design of steel structures — Part 1 — 5: Plated structural elements.
EN 1995-1-1:2004/AC:2006/A1:2008/A2:2014 Eurocode 5:
Design of timber structures — Part 1 — 1: General — Common rules and rules for buildings.
EN 1995-2:2004 Eurocode 5: Design of timber structures — Part 2: Bridges.
EN 14080:2013 Timber structures — Glued laminated timber and glued solid timber
— Requirements.
EN 14358:2006 Timber structures — Calculation of characteristic 5-percentile values and
acceptance criteria for a sample.
EN ISO 8560:1986 (1999) Technical drawings — Construction drawings — Representation of modular sizes,
lines and grids.
Non-liability
By using the content of Design of timber structures - Volume 1, you agree to the terms of use below.
The authors and reviewers have been responsible for interpreting building regulations, research
papers, industrial documents and so on, with a view to conveying current design practices.
The material presented is only intended to provide guidance; the final design responsibility lies with
the structural engineer or designer.
All information is provided in its present state and without any form of guarantee, to the extent
permitted by prevailing legislation. Although the publisher has taken all reasonable steps to try and
provide reliable information in Design of timber structures - Volume 1, the publisher does not guarantee
that the content is free from errors, mistakes and/or omissions, or that the content is up to date and
relevant to the user’s needs.
The publisher, Swedish Forest Industries Federation, also provides no guarantees for any outcomes
arising from use of the information contained in Design of timber structures - Volume 1. All use of
information in Design of timber structures - Volume 1 is the user’s own responsibility and is done at
the user’s risk.
Swedish Forest Industries Federation does not accept liability for any damage or injury that may arise
due to the content of Design of timber structures - Volume 1.
Copyright
Swedish Forest Industries Federation owns the rights to the content
Design of timber structures - Volume 1.
The content is protected under copyright law. Misuse will be subject to a penalty.
Copying of the content is prohibited.
Language reviewer
Patrice Godonou — Swedish Wood
Project managers
Johan Frobel — Swedish Wood Illustrations
Patrice Godonou — Swedish Wood Vendela Martinac — Thelander Arkitektur & Design AB
Cornelia Thelander — Thelander Arkitektur & Design AB
Editor
Patrice Godonou — Swedish Wood Photo credits
Tommy Andersson, page 182
Authors Johan Ardefors, page 208 (top)
Chapter 1 Per Bergkvist, page 19
Robert Kliger — Chalmers University of Technology, Anders Bjornfot, page 217 (top)
Structural Engineering, Goteborg, Sweden Patrick Degerman, page 47, 200, 208 (middle and lower)
Bertil Enquist, page 29
Chapter 2 Ake E:son Lindman, page 4, 9, 15, 43, 53, 57 (right), 116, 198
Marie Johansson — LNU, Linneus University, Gremo, page 25 (right)
Department of Building Technology, Vaxjo, Sweden Anders Gustafsson, page 217 (lower)
Thomas Harrysson, page 212
Chapter 3 Soren Hakanlind, page 209 (top and lower right), 119,
Roberto Crocetti and Annika Martensson — 189, 191
Faculty of Engineering, LTH, Lund University, Marie Johansson, page 24
Division of Building Science, Lund, Sweden Kerstin Jonsson, page 26, 27
Martinson Group AB, page 240
Chapter 4 and 8 Rottne Industri AB, page 25 (left)
Helena Lidelow — LTU, Lulea University of Technology, Setra Travaror AB, page 11
Department of Civil, Environmental and Natural Stora Enso, page 202
Resources Engineering, Lulea, Sweden Gosta Wendelius, page 209 (lower left)