You are on page 1of 123

Lecture Notes

Introduction to Timber
Content
Content ....................................................................................................................... 1
Part 1: GENERAL ...................................................................................................... 2
1 Introduction .......................................................................................................... 2
2 Use of wood ......................................................................................................... 4
3 Coniferous and deciduous wood [3] ..................................................................... 6
4 Dendrochronology ...............................................................................................10
5 Wood properties ..................................................................................................12
5.1 Density .........................................................................................................12
5.2 Wood Moisture Content ................................................................................13
5.3 Shrinkage and Swelling ................................................................................15
5.4 Fungal degradation.......................................................................................16
5.5 Strength and stiffness ...................................................................................16
5.6 Load duration................................................................................................16
5.7 Load duration classes...................................................................................17
5.8 Volume effects ..............................................................................................17
5.9 Heat-properties .............................................................................................19
5.10 Electrical conduction.....................................................................................19
5.11 Durability ......................................................................................................20
5.12 Preservation .................................................................................................21
5.13 Modification ..................................................................................................22
5.14 Summary ......................................................................................................22
6 Structural design calculations .............................................................................23
6.1 Strength analysis (Safety) ............................................................................23
6.2 Stiffness analysis (Serviceability) .................................................................27
7 Grading: wood quality and strength....................................................................29
7.1 Quality classes and strength classes............................................................29
7.2 Quality marking.............................................................................................33
8 Wood products ....................................................................................................36
9 Literature .............................................................................................................37
Part 2: DESIGN OF WOODEN LOAD-BEARING STRUCTURES ...........................38
10 Introduction ......................................................................................................38
11 Rules of thumb for girders and columns ..........................................................40
11.1 Rules of thumb for girders ............................................................................40
11.2 Rules of thumb for columns ..........................................................................45
12 Rules of thumb for three-hinge-frames ............................................................46
13 Structural detailing ...........................................................................................48
14 Stability ............................................................................................................54
15 Literature..........................................................................................................58
Annex 1: Calculation example: Unity Checks of the stresses in a single span beam59
Annex 2: Tables Eurocode 5 ....................................................................................62
Introduction to Timber Structures 1
Annex 3: Floor beams: ratio span versus beam height (rules of thumb) ..................66
Annex 4: Rules of thumb for Timber Constructions ..................................................69
Annex 5: Standard timber sizes................................................................................74

Introduction to Timber Structures 2


Part 1: GENERAL

1 Introduction

For centuries wood has been an important construction material. After the
development of "modern" materials like concrete and steel, a steady decline in the
amount of wood as construction material occurred. This is illustrated in Figure 1.1
[1].

Figure 1.1 Development of wood in the construction as a


percentage of the total material usage.

Since a number of years the application of wood in constructions is increasing. The


material is "rediscovered". Especially the visual qualities of the material, whatever
that may be, contributes to this increase.
Also other positive qualities such as :
• the high strength and stiffness parallel to the grain in relation to the mass,
• the high fire resistance,
• “free form” production possibilities,
• chemical resistance, the natural durability,
• the very low energy use in production
• and very low emissions of so-called greenhouse gases (e.g. CO2) contribute.

The low natural durability of different wood species is often seen as a negative
property. However, the natural durability is no indicator for sustainability of structures
realized with certain wood species.
Wood constructions realized with low durable wood species still function after
hundreds of years because the conditions for damaging mechanisms (fungi –
insects – bacteria) are absent or very low in many circumstances.
Introduction to Timber Structures 3
The variation in wood appearance is very large. Even within one tree significant
differences occur. This results in possible fascinating designs. Also with regard to
the structural properties big variation can be expected. Also in this case: every piece
of wood is different. Result is that considerable efforts need to be made to get a
well-defined impression of the properties. Hundreds, if not thousands, of wood
species exist. Only a limited number, ca. 45, are classified in so-called strength
classes, so that these species can be used structurally in a sound, well defined way.
However, for most of these species all the wood is classified in only one strength
class limiting the efficient use.

Introduction to Timber Structures 4


Wood is a natural material. It can be applied directly as raw material (round wood);
e.g. as foundation pile, which outnumbers all other raw material usage. A rough
estimate indicates, that in the Netherlands, especially in the western part (Holland)
about 25 million tree trunks used as foundation piles support all kind of structures.
Recent data provided by suppliers of wooden foundation piles, indicate that there is
still a (small) market for wooden piles in Netherlands and Flanders.

Wood can be cut (sawn) into all kind of elements (girders, columns, boards, laths,
etc). Sawn elements can be combined into so-called composite elements, with glue
(glulam) and/or mechanically (wire nails, bolts, etc.). Veneer, usually obtained by
peeling the tree, is glued on less attractive wood species to obtain plate material for
furniture manufacturing or finishing panels. Veneer layers can be glued together to
obtain, plywood, a sheet material, or LVL (laminated veneer lumber). Thicker boards
glued together crosswise result in so-called CLT (cross laminated timber).
From waste, from sawdust to chips other sheet materials like particle board,
composed plate OSB (oriented strand board), MDF (medium density fibre board) and
HDF (high density fibre board) are manufactured. Generally the structural properties
reduce with increasing amount of glue needed (e.g. MDF needs relatively much glue
and plywood relatively little). All these products are used in the construction, often as
finishing (embellishing, fire resistance, etc.). In a number of cases structurally for
which plywood, CLT, LVL, fibre board, OSB and particle board are most suited.

Introduction to Timber Structures 5


2 Use of wood
Timber and timber products are obtained by cutting trees and processing the raw
material in some way. Figure 2.1 shows the use of this wood, averaged over the
entire world.

Figure 2.1 Wood use (worldwide) [2].

Figure 2.1 shows, that more than half of the felled trees is used as firewood. In
Western Europe, North America and Japan the ratio is completely different. There is
the portion firewood relatively small (can be ignored). For the Netherlands the use of
wood is shown in Figure 2.2.

Figure 2.2 Wood use (The Netherlands) [2].

The use of the forests, the suppliers of wood, differs. These different
purposes do not always strengthen each other (these purposes compete).
Different purposes are:
• Human shelter (especially in certain areas in Africa, Asia and South
America) and animal habitat.
• Recreation (especially in the so-called Western world).
• Supplier of food, medicine, etc.
• Supplier of wood for all kinds of applications.
• A positive contribution to the so-called CO2 (carbon dioxide) balance.

Worldwide about 0,5 m3 of wood is used per person per year. In Netherlands this
is approx. 1 m3, in Japan and the United States even slightly higher.
Introduction to Timber Structures 6
Socially, ecologically and economically, it is becoming less accepted, that the used
wood contributes to the negative effects of deforestation of the Earth. It is more and
more expected that the wood used is coming out of sustainably managed forests,
aimed at forest conservation and at positive socio- economic development for all
involved.
most used (Dutch names between brackets)

Deciduous wood species are mostly used in furniture, facade carpentry (window
frames, doors, etc.) and civil engineering structures (bollards, scaffolding, etc.).
This wood is predominantly originated from tropical forests.

3 Coniferous and deciduous wood [3]


There is a clear difference in structure, properties and growth area between
coniferous and deciduous wood species. In general coniferous wood is softer,
lighter and less strong than deciduous wood. Also, generally, the natural
durability of coniferous wood is less compared to deciduous wood.

Most coniferous wood species grow between 70 and 50 latitude as a belt bordering
the Arctic across North America, Europe and Asia. Also in the southern hemisphere
forests with coniferous wood species can be found (e.g. in in New Zealand, South
Africa and Argentina-Chile, where large areas are forested with so-called Radiata
Pine trees).

Tropical forests (in which deciduous trees are found) are present in the rest of the
world: globally in the temperate, subtropical and tropical zones between the 50
latitude north latitude and 50 South latitude. Figure 3.1 shows the global growth
areas.

Introduction to Timber Structures 7


Figure 3.1 Growth areas (something more than 25% of the land
surface is forested worldwide) [3].

Deciduous trees respond strongly to the changing of the seasons. The food intake in
the root structure occur especially in spring time resulting in a relative high growth
rate (creating so-called early wood) and leave creating, through which moisture
evaporates. Summer growth conditions for the tree are more hazardous resulting in
slowing down of the growth rate and leave creating, which almost fully stops in
autumn and winter periods. The wood created in summer, autumn and winter
periods, the so- called “late wood”, is denser than the wood created in springtime.
The “late wood” is visualized by the so-called annual rings.

The wood structure of deciduous wood species is different from that of


coniferous wood species. There are also similarities. Both are shortly discussed
with the help of figures 3.2, 3.3 and 3.4.

Introduction to Timber Structures 8


Figure 3.2 Tree trunk cross section and three dimensional trunk views.

Fibre structure
A tree grows in two directions, longitudinally and radially (height and in thickness
directions). Actually, these cells stretch themselves. The thickness growth in the
stem, the roots and the branches takes place in the cambium, a single cell layer
between the bark and the sapwood. Growth rings are formed.

The cambium cells remain active


throughout the life of the tree. By cell
division to the outside the inner bark
(also called soft bark) is formed. By cell
division to the inside, wood cells
(sapwood) are formed. Usually, the
number of cells formed for the inner
bark is much smaller than the number
of wood cells (a tree contains more
Figure 3.3 Coniferous wood (Pine)
wood than bark).

Nutrition from the roots to the tree crown takes place through the outer parts
(inner bark to sapwood) of the trunk: in the sapwood the rising flow takes place,
the inner bark the descend flow. Horizontal moisture transport is possible via the
rays.

After some years, the middle section no longer contributes to the nutrition transport.
This part, the heartwood, "dies" after the cells present herein may be filled with

Introduction to Timber Structures 9


natural chemicals to enlarge natural durability, after which in most species these
cells are closed (for durability reasons too).

Introduction to Timber Structures 10


In deciduous wood species the organisation of the fibres and fibre groups (tissues)
looks much less uniform than in coniferous wood species. In deciduous tubes, called
vessels, develop if cells grow together in length direction. For deciduous wood, these
vessels are an important feature. Softwood lacks these vessels. In coniferous wood
the moisture transport (nutrition transport) from one cell to the other takes place
through (small) lockable openings, called bordered pits (pits that connect cells), see
Figure 3.4.

Figure 3.4 Difference in structure between coniferous and deciduous


wood [3].

All important wood properties such as strength, stiffness and shrinkage / swell
behaviour are explained on the basis of the chemical structure and the anatomy, or
the structure of the cells and the properties of the cell wall material. If the cell wall
material degenerates due to e.g. wood degradation or fire, the material and thus the
mass and strength properties reduce.

Because wood is a natural product, growth “failures” and imperfections like knots
and growth disturbances (due to geometrical imperfections) can be expected. The
wood grain is not completely straight resulting in strength and stiffness reductions.

Introduction to Timber Structures 11


Wood is characterized by its fibre direction (longitudinal direction). Due to this typical
structure, the wood anatomy, physical and mechanical properties vary widely in
different directions (e.g. the material strength is higher parallel to the fibre direction
than perpendicular to the fibre direction).

Introduction to Timber Structures 12


When observing wood with help of an electron microscope, the cells can be seen as
an integrated bundle of fibres. These can be compared to vertical tubes. Under the
microscope also the annual rings can be seen. The fibres (tubes) can be loaded in
different directions. The strength and stiffness properties vary in different directions.
This is meant when talking about the anisotropic nature of wood. In axial direction, or
parallel to the fibre direction, wood very strong, both in tension as in compression.
Perpendicular to the wood fibre, on the other hand, wood is less strong and stiff.
Compression perpendicular to the fibre direction may lead to fibre crushing, which
does not lead to failure (and is therefore a safe “failure mechanism). However,
tension perpendicular to the fibre direction, wood is weak. Tension perpendicular can
lead to brittle failure which must be avoided (unsafe failure mechanism). See also
Figure 3.5.

The cell walls thickness and the size of the


void cavities varies per wood species. The
density of the cell wall material itself is
independent of the type of wood and is
approximately 1550 kg/m³; a high density
of wood means that the cell walls are thick
and robust.

Figure 3.5 Directional sensitivity of The density of wood is the mass of 1 m³ of


wood.
wood at specific moisture content, usually
12% or 15% (relative to the dry mass, see
chapter 5). Generally
heavier wood species show higher values
for the
mechanical properties than lighter wood species. That is also logical. A high density
means a relatively high amount of wood fibre material per mm² cross-section and
consequently the strength properties (tensile, pressure and bending) parallel to the
fibre directions show relatively high values.

The strength properties generally exhibit a slightly larger variation in softwood than in
deciduous wood, due to the difference in the presence of knots. Generally deciduous
wood is also heavier than softwood, which is reflected in the strength properties. The
differences in wood structure and the dependence of the strength and stiffness on the
fibre direction is clearly shown in the strength classes of wood (tables 6.1 and 7.1).

4 Dendrochronology
Dendrochronology, or tree ring research, is the scientific discipline that deals with
Introduction to Timber Structures 13
dating of wooden objects or archaeological findings on the basis of recognizable
annual rings in the objects (growth rings). Large parts of the world exhibit seasons.
The largest growth occurs in the spring, the smallest in the winter. As a result, the
spring wood (early wood) with large wide cells distinguishes itself, see Figure 3.4,
from the late wood (formed in summer, autumn and winter). This results in a
"circular" lines pattern, also known as annual rings. This enables the age
determination of a tree, by counting the number of rings between the edge and the
center (pith). This can easily be done after the tree is cut down. For still standing
trees this can also be done after retaining a wood cylinder form the tree using a
special wood drill, designed for this purpose (increment drill, which is a hollow drill).

Introduction to Timber Structures 14


Because trees of the same species from the same area are exposed to the same
climatic conditions exhibit the same pattern in the same growth years; broad annual
ring pattern (generally a sign of good growth conditions), narrow annual ring pattern
or abnormalities on the growth ring pattern.
By comparing ring patterns of trees from the same growth area with an overlap in
time (two time periods with an overlap), these time periods can be linked to one
period. By linking several periods together, the age of wooden objects found can be
determined which is important for the identification of remains of old construction
and/or archaeological findings.
The annual ring patterns can be captured by studying a slice of wood under a
microscope and by measuring accurately the width (and disturbance) of each annual
ring. Based on these measurements, for several regions of provenance dating
calendars are developed.

Thus, dendrochronology is not limited to living, or recently felled, trees. With a piece
of (archaeological) wood containing older unknown annual ring information in
addition to known annual ring information the dating calendar can be expanded.
Figure 4.1 shows this principle based on a tree cut in 1973 whose ring pattern
perfectly suits the growth ring patterns in older wood.

Figure 4.1 Use of dendrochronology.

It is essential to realise that the dating calendars are attached to a combination of


growth area and wood species. Consequently, the correct dating calendars must
be used. If this is done correctly, information on the growth areas of the wood
used can be obtained. Based on these studies it is shown that the oak for the
Dutch ships during the Dutch Golden Age was obtained from the Baltic Area
Introduction to Timber Structures 15
(North Estonia, Latvia, Lithuania, Poland, Germany).

Introduction to Timber Structures 16


5 Wood properties
The following items are briefly discussed in this chapter
• density
• wood moisture content
o shrinkage and swelling
o durability; biological resistance
o strength and stiffness
• load duration
o strength and stiffness
• volume effects
o strength and stiffness
• heat properties
o conduction coefficient
o heat capacity
o expansion coefficient
• electric properties
o conduction
• chemical characteristics
o chemical degradation
o chemical resistance

5.1 Density
The density of wood varies from ca. 150 kg/m3 (balsa) until ca. 1230 kg/m3 (lignum
vitae). The density is calculated with formula (5.1).
 =
 m
[kg/m3] (5.1)

V

With density at the wood moisture content 




m mass at the wood moisture content 
V volume when the wood moisture content 

It is usual to determine the density at a wood moisture content (  ) of  = 12 %.

The mass of the solid material in the cell wall is, at  = 12 %, approximately 1550
kg/m3 for every wood species. The proportion of air, the hollow spaces (void
cavities), is the reason for the large differences in density between the different wood

Introduction to Timber Structures 17


species. The amount of void space relative to the material without void cavities is
expressed in the pores share [%] that is calculated using formula (5.2).

p = 1001 − 12 [%] (5.2)


 
 1550 

Introduction to Timber Structures 18


For three wood species the pore share is calculated in table 5.1.

Table 5.1. Pore share for three wood species.


wood species density [kg/m3] Pore share [%]
1230
Lignum vitae (pokhout) 1230 100 − 100 = 21
1550
440
Spruce 440 100 − 100 = 72
1550
150
Balsa 150 100 − 100 = 90
1550

5.2 Wood Moisture Content


The wood moisture content is calculated according to formula 5.3.
− m
= 
m
=0% 100 [%] (5.3)
m =0%

Formula (5.3) shows the definition of the wood moisture content: the mass of the
water in the wood, expressed as a percentage of the mass of the dry wood. Note
that wood moisture content values higher than 100% is realistic. For example: the
void cavities in spruce occupy, see table 5.1, 72% of the total (wood) volume. If
these void cavities are filled with water the wood weighs ca. 400 + 720 =
1120 − 400
1120 kg/m3 and the wood moisture content   100 = 180 % (+12%) .
400
Moisture in wood is partly bound and partly free. Bound water is chemically bound
to the cell walls. Free water is located in the hollow spaces (void). If all positions
where moisture can be bound to the cell walls are occupied, the wood has reached
the so-called saturation point (FSP: Fibre Saturation Point); see also Figure 5.3.

The wood moisture content is governed by climatic conditions, described with the
relative humidity (RH) and temperature, see Figure 5.1.

Introduction to Timber Structures 19


Figure 5.1 Relationship between the wood moisture content
(equilibrium moisture content) and climatic conditions
(Temperature and Relative air humidity).

Introduction to Timber Structures 20


The wood moisture content that sets itself in certain climatic conditions, mainly
determined by the relative air humidity, is called equilibrium moisture content.
With regard to this equilibrium moisture content three climate classes are defined as
can be seen in figure 5.1.
• Climate class 1: the equilibrium moisture content is expected not to exceed  =
18 %.
• Climate class 2: the equilibrium moisture content is expected not to exceed  =
20 %.
• Climate class 3: the equilibrium moisture content is dominantly higher than
20% (   20 %). These climate classes are defined in table 5.2 in a bit more detail.

Table 5.2. Climate classes.


Climate ave Description
class rage
[%]
1 12 Standard indoor conditions
2 20 Outdoor, covered structures
3 >20 - poorly ventilated spaces (indoor)
- fully exposed to outdoor conditions (not
covered)
- structures in and underneath water

If the climatic conditions change, the wood moisture content will also change. The
change rate of the wood moisture content depends on the wood species, the
surface area of end grain cut (moisture exchange with the environment through end
cut surfaces(parallel to the grain) is much faster than perpendicular to the grain)
and dimensions of the wood. In any case there is a delay between the changes in
climatic conditions and setting the associated wood moisture content (preferably the
equilibrium moisture content). This is illustrated in Figure 5.2; the graph shown in
figure 5.2 is called hysteresis.

Introduction to Timber Structures 21


Most wood species show a fresh wood
moisture content of 50-60% (moisture
content immediately after felling). There
are, of course, exceptions such as the
Poplar, with a fresh wood moisture
content above 100% (a result is that the
water from the wood flows out of the
wood after felling).

Wood under water is completely


saturated and most hollow spaces are
Figure 5.2 Hysteresis [4]. completely filled with water.
adsorption: increasing wood moisture
content desorption: decreasing wood The bound water influences the wood
moisture content properties up to a high extend. The
properties are hardly affected by
moisture content changes above Fibre
Saturation Point; see figure 5.3.

Introduction to Timber Structures 22


Figure 5.3 Effects of the wood moisture content on wood properties.

5.3 Shrinkage and Swelling


The dimensions of wood react to changing wood moisture content:
• Wood shrinks when the wood moisture content reduces.
• Wood swells when the wood moisture content increases.
This applies, however, only if the wood moisture content is less than the FSP (Fibre
Saturation Point).

Wood shrinks and swells


negligible parallel to the wood
fibre direction. Shrinking and
swelling in perpendicular to the
grain directions cannot be
neglected. From the
perpendicular to the grain
directions the dimensional
changes are most expressed in
tangential direction being about
twice as big as in radial direction,
see Figure 5.4.

Figure 5.4 Shrinkage and swelling in


perpendicular to the grain directions
(radial and tangential) [4].

Introduction to Timber Structures 23


5.4 Fungal degradation
Wood is a natural (living) material, which serves as nutrition for many mechanisms
resulting in degradation of the wood (structural) properties. One of the best-known
effects of this degradation due to fungi attack resulting in wood rotting. The wood
rotting fungi can for sure not live below a certain wood moisture content. A safe
upper limit value is wood moisture content of about 21%: fungi do not develop in
wood when the wood moisture content does not exceed 21% (fungi develops for
sure for wood moisture contents exceeding the Fibre Saturation Point.

5.5 Strength and stiffness


Dry wood is stronger and stiffer than wet wood. In general, this applies,
however, only at wood moisture content below the Fibre Saturation Point (see
figure 5.3).

From the above it can be concluded, that wood moisture content values exceeding
the Fibre Saturation Point, result in filling up the void spaces (free water), and
almost exclusively causing the density to increase.

5.6 Load duration


The mechanical wood properties (strength-and stiffness) respond to load duration:
long term loading result in increased deformations. This phenomenon is called
creep (increasing deformation under constant load), shown in figure 5.5.

Figure 5.5 Effect of the load duration (and wood moisture content) on
the deformation of a permanently loaded beam.

On the other hand, the stresses reduce in time when an element is deformed with
a constant value; this phenomenon is called relaxation (decreasing internal
stresses under constant deformation).
The strength of wood is reduced for long term loaded elements compared to the
strength of short term loaded elements. This is illustrated in Figure 5.6.
Introduction to Timber Structures 24
Figure 5.6 Effect of load duration on strength.

The load duration effects on strength are described on experimentally-based


models. One of these models, the so-called Madison Curve [5], is indicated in
Figure 5.6. This is a so-called regression equation based on many experiments
(with which only the average values are accessed). The Madison curve is
important, because the modification factors kmod , involved in construction
calculations, described in formula (6.1), are based on this model.

5.7 Load duration classes


The load duration of loads differ (e.g. the load duration of permanent loading is
much longer than of short term variable loading). Therefore, the loads usually
applied are classified in so-called load duration classes. Table 5.3 shows the load
duration classes defined in Eurocode 5 [9].

Table 5.3. Load duration classes.


Load duration class Cumulative duration of the Examples
characteristic load
Permanent Longer than 10 years Dead load
Long 6 months - 10 years Storage
Medium-Long 1 week - 6 months Life loads on floors
Short Less than 1 week Snow, wind (The
Netherlands)
Instantaneous Accidental load, wind
(Belgium)

5.8 Volume effects


Volume effects are perhaps best to understand on the basis of a loaded chain. If one
Introduction to Timber Structures 25
link in the chain breaks, the load carrying capacity of the complete chain (the whole
system) has disappeared. Actually this illustrates so-called brittle failure.

Consider the chains in Figure 5.7. The individual links have certain strength. Due
to variation in individual link properties, the "links" are of wood, the individual links
vary in strength. This strength

Introduction to Timber Structures 26


variation is also shown in Figure 5.7. The chance that besides strong links also (very)
weak links are present increases with increasing chain length: in figure 5.7 chain (a)
is most probably less strong than the shorter chain (b).

In analogy the probability on a weak spot in a large timber element is greater than in
a small timber element. The chain represents so-called brittle failure modes (no
redistribution possible: if one link breaks, the full system fails). For practical
calculations, this volume effect is translated into a volume / height factor. This factor
is therefore only applicable to those material properties showing brittle failure modes.
Wood in tension and/or bending show brittle failures and consequently the factor
applies to these material properties. On the other hand, wood under compression
shows tough failure behaviour and consequently the volume Factor (or height factor)
is not applicable to compression.

Figure 5.7 Brittle Fracture (virtually linear behaviour until failure).

The values for the volume / height factors are derived from the so-called
Weibull distribution (a probability distribution), displayed with formula (5.4).

Probability [ 1 = f for a given Volume V1 and given probability [ 0 = f for a given


Volume V0 ] (5.4)
In 1 = tensile stress level in volume V1 [N/mm2]
which
0 = tensile stress level in (reference volume)
volume V0
f = tensile strength [N/mm2]

The tensile strength corresponds to a volume V0 = 0.01 m3 which is subjected to a


uniform tension
stress. The is so-called reference volume. Elaboration of formula (5.4) results
volume V0 in, see [10],
formula (5.5).
V 0,2
Introduction to Timber Structures 27
 max;d
 0   hV  f [N/mm2] (5.5)
d 0,2 d
 1 
V  ((x, y, z))dV 
 Vh h 

In practice, formula (5.4) is for tension and bending parallel to the fibre direction
simplified to a height
factor kh and length factor kl . The length factor is exclusively used for laminated
veneer lumber
(LVL); see Chapter 7.

Introduction to Timber Structures 28


The fact, that, at least for bending, the volume factor can be reduced to a one-
dimensional height factor is probably best to understand if one realizes that the span
(length), the beam height and beam width are linked to each other. The beam height
for beams can approximately be set to:

span
for roof beams
20 to
25
span
15 to for floor beams
20

5.9 Heat-properties
Heat conduction, heat capacity and expansion due to temperature increase are
defined.
 W 
The heat conduction is expressed with the conduction coefficient  . Wood is a
material with

m  K 
different properties in different directions. In radial and tangential direction both
perpendicular to the grain directions, the values for  globally equal. The value parallel
to the wood fibre is 2 to 3 times as large. Wood has, relative to many other materials,
a low thermal conductivity. This makes wood a suitable material in heat-insulating
constructions. See table 5.5 in which some values for  are given.

The  energie 
Joule heat
 = capacity is expressed in .

 m3  K   m3  Kelvin 
 Joule 
3
Dry wood has a specific heat of ca.  . With a mass of 500 kg/m , this leads to a
1880  heat
 kg 
 Joule 
. Stone has a specific heat of ca. 840 
3 Joule
capacity of 5001880 = 94010
 . At a mass of
 m3  K  m3  K
 
 Joule  . From this it can be
1800 kg/m3, this leads to a heat capacity of  m3  K 
1800840  1500103
concluded that a wooden wall heats up considerably faster than a stone wall. The
wooden wall cools down, however, also significantly faster. From this it can be

Introduction to Timber Structures 29


concluded, that for energy reduction spaces which are not constantly heated can
better be realised in wood than in stone.

Wood tends to expand when temperature increases. At the same time the wood
moisture content reduces resulting in shrinkage. The effect of shrinkage due to
moisture decrease is much larger than the temperature expansion and
consequently thermal expansion is seldom regarded.
For European coniferous wood the following values for the coefficient of thermal
expansion can be
used: 0  410−6
; ;
 2010 −6  2010
−6 90;radial
90;tangential

5.10 Electrical conduction


Dry wood conducts electricity poorly. Wet wood on the other hand, conducts
electricity well. The extent to which electricity conducts depends, below Fibre
Saturation Point, on the wood moisture content.
Based to this property electrical moisture content measurement devices have been
developed; figure
5.8 shows one.

Introduction to Timber Structures 30


So-called calibration graphs are determined for different wood species, which must
be set before the measurements can start.

Voltage difference between


pegs initiating an electric
current of which the value
depends on the wood
moisture content.

The pegs are driven into the


wood, perpendicular to the
fibre direction, for about 5 to
30 mm.
Figure 5.8 Electrical moisture measurement device.

5.11 Durability
The wood moisture content has a big impact on the durability of timber structures.
This is illustrated in Figure 5.3: provided that the wood moisture content is ca. 21%,
no fungi (causes of wood rotting) develop. Durable detailing and construction is
based on reducing the wood moisture content to below 21%.

Introduction to Timber Structures 31


There are big differences in the so-
called natural durability between
different wood species. General
statements in the practice of "common
deciduous wood is more durable than
softwood" and "tropical (hard) wood
species (these are deciduous
species) are much more durable than
the most commonly used uses
coniferous wood species" are to a
certain extent true. These statements
need, however, nuances which can
be understood by studying the tree
cross section from Figure 5.9.

Introduction to Timber Structures 32


Figure 5.9 Tree (trunk) cross Roughly spoken, a tree consists out of
section
dead area, the heartwood, and a living
area, the
sapwood, cambium and inner bark (the
outer

bark protects the tree, e.g. against forest fires, an consists out of dead material as
well).
Before on the border of heartwood and sapwood wood cells are added to the
heartwood, many wood species add components to these cells (e.g. natural toxins),
which increase the resistance against fungal degradation. In addition, the cells are
closed (in coniferous wood species the border pits, see Figure 3.4, are closed). After
the addition of the cells to the (dead) heartwood, these cells are no longer active and
the tree itself is no longer capable to protect these cells. Some wood species, e.g.
beech, do not protect the cells before adding to the dead heartwood which explains
that the heartwood of these wood species can be destroyed by fungi completely.

The living part (mainly sapwood) is protected by the tree itself. At the moment the
tree is cut, this part is hardly affected by fungi, insects, etc. On the other hand, the
tree did not take precautions to protect this part. In other words, the sapwood is not
protected like the heartwood. Consequently, the sapwood of each type of wood has
a rather low natural durability. The variation in natural durability between different
wood species is therefore only true for the heartwood. Consequently, the durability
classes given in table 5.4 only reflect the heartwood.

The heartwood of different wood species is classified in a so-called durability


class based on experimental research, the so-called "graveyard” test.
.
Five durability classes are distinguished, see table 5.4.

Table 5.4. Durability classes (natural durability).


class Time [years] in ground Wood Species
contact without fungi HEARTWOOD
attack
I ≥ 25 Azobé, Afzelia, Bilinga, Cumaru, Iroko,
Piquia, Masseranduba, Teak
II ≥ 15 European oak, Basralocus, sweet chestnut,
mahogany, Merbau, Robinia, Wengé,
Western Red Cedar
III ≥ 10 Larch, Douglas Fir, white American oak, Meranti,
Pitch
Pine
IV ≥5 Pine, American oak, Hemlock, spruce,
V <5 Beech, Poplar, Birch, Radiata Pine

Notes: (1) durability is always guaranteed when the wood moisture content does not
exceed 21%. In that case all wood species, regardless the natural durability,
Introduction to Timber Structures 33
can be applied.
(2) under certain conditions wood from durability class III can be applied in
an unprotected outside environment (climate class 3), see table 5.2. From
table 5.4 it follows that for this application the heartwood of deciduous wood
species (tropical) is most suitable. For (almost) vertical elements however,
the heartwood of Larch and Douglas Fir (durability class III) are suitable as
well in these conditions.

5.12 Preservation
To prevent the wood from fungi attack it can be treated with toxic substances.
This is called wood preservation. Two of the most well known processes are the
so-called "waterborne preservatives",

Introduction to Timber Structures 34


very suitable for pine, and “coal tar creosote”. During the waterborne preservatives
process copper- chromium-arsenic compounds, solved in water, are pressed into
the wood. During the “coal tar creosote” process liquids based on hydrocarbons
are pressed into the wood. Both processes are realized with vacuum-pressure
cycles.
For environmental reasons and due to social acceptance, the legal and social
opportunities for wood preservation with toxic compounds are increasingly
restricted in recent years.

5.13 Modification
Due to the fact that preservation with toxic compounds is increasingly restricted in
recent years, new environmental friendly alternatives have been developed. The so-
called modification techniques, both thermal as chemical, are well upgraded to
industrial scale and used more and more frequently.

5.14 Summary
Table 5.5 shows a number of wood properties discussed in this chapter. A much
more extensive table can be found in the “Houtvademecum” [6] (in Dutch).

Table 5.5. Some wood properties at a glance [6].


shrinkage (at 6 ≤ ω ≤

thermal expansion [mm /

durability class (heartwood)


specific heat [J / kg · K]
moisture content ω [%]
density [kg/m3]

species
λ [W / m·K]*

remark
20)

K]

rad. tang. rad. tang.


European beech 720 1 0.17 0.36 V deciduou
2 s
European oak 720 1 0.16 0.26 0.1 II deciduou
5 8 s
Silver Fir 0.12 0.27 0.1 IV coniferou
1 s
Scotch pine 460 1 0.15 0.30 0.1 IV coniferou
2 3 s
20·10-6

35·10-6

Oregon Pine 530 1 0.18 0.31 0.1 III coniferou


1880

(Douglas) 4 3 s
Edible Chestnut 540 1 II deciduou
2 s
Larch 590 1 0.1 III coniferou
Introduction to Timber Structures 35
2 3 s
Norway Spruce 440 1 0.14 0.26 0.1 IV coniferou
2 1 s
Steel 785 50 12·106
0

*according to ISO 10077-2 for window frames

Introduction to Timber Structures 36


6 Structural design calculations

6.1 Strength analysis (Safety)


For structural design calculations strength and stiffness properties of materials
used are essential. Analyses related to safety are carried out to predict the ultimate
load carrying capacity (failure level); for these calculations so-called ultimate limit
states (ULS) are recognized. For this purpose so-called design values have to be
defined for which formula (6.1) is used.
f
f = k k  k (6.1)
k 
d mod
h l
m

Wit fd design value of the strength-property


h:
fk [N/mm2] characteristic value of the
km strength-property [N/mm2]
od modification factor, mainly determined by load duration. To a lesser
extent determined
by climatic conditions in which the structure is located (to which the
kh ,
wood moisture content is related).
kl
factors, taking structural element dimensions into account (see
volume effects in Chapter 5).

Characteristic value of a strength property


The characteristic strength value depends on the type of wood and the wood quality.
Based on these two identifications the wood is classified into so-called strength
classes. The strength classes are, with characteristic strength values, given in EN
338 [7] for sawn timber and in EN 1194 [8] for glued laminated timber. Table 6.1 is a
short version of the given tables in EN 338.

Table 6.1. Strength classes [7].


Strength class C18 C24 D30 D40 D50 D70 GL24
h
Ultima fm,k N/mm 18 24 30 40 50 70 24
2
te
Limit ft ,0,k N/mm
2
10 14.5 18 24 30 42 19.2
States
ft ,90,k N/mm
2
0.4 0.4 0.6 0.6 0.6 0.6 0.5
(ULS)
fc,0,k N/mm 18 21 24 27 30 36 24
Introduction to Timber Structures 37
2

fc,90,k N/mm
2
2.2 2.5 5.3 5.5 6.2 12.0 2.5

fv ,k N/mm 3.4 4.0 3.9 4.2 4.5 5.0 3.5


2
k kg/m 320 350 530 550 620 800 385
3
Em,0,k N/mm
2 6,00 7,400 9,200 10,90 11,80 16,80 9,600
0 0 0 0
Seviceabil Em,0,m N/mm 9,00 11,00 11,00 13,00 14,00 20,00 11,50
2
ity Limit ean 0 0 0 0 0 0 0
Em,90, N/mm 300 370 730 870 930 1,330 300
States 2
(SLS) mean
Gmean N/mm 560 690 690 810 880 1,250 650
2

Note: more detailed information follows in Chapter 7.

Introduction to Timber Structures 38


Material factor  m
The material factor  m , see formula (6.1), depends on the accuracy with which the
characteristic values of the material properties can be determined. For sawn wood
this is depending on the accuracy with which the wood can be classified into strength
classes.
Industrial manufactured products such as glued laminated wood, plywood and
Laminated Veneer Lumber (LVL) exhibit less variation in strength and stiffness
properties than sawn wood.
Consequently, the characteristic values of industrial manufactured products can be
determined more accurate. For this reason, the material factors for industrial
manufactures products are smaller. Table
6.2shows the in EN 1995-1-1 (Eurocode 5) [9] given material factors.

Table 6.2. Material factors  m .


materi Material factor
al m
sawn timber 1.30
glued laminated 1.25
wood
LVL, plywood, OSB 1.20
connections 1.30
metal plate 1.25
connectors

For structural calculations using the accidental load combinations, e.g. in seismic
design, the material factors all reduce to  m = 1.0 . Also for calculations in
Serviceability Limit States  m = 1.0 .

Modification factor kmod


The characteristic strength and stiffness properties, given in table 6.1, are based
on the short-term strength values, obtained by testing in which failure is obtained in
roughly 300 +/- 200 seconds, at a wood moisture content of 12% (climate class 1
according to table 5.2).
If the load duration differs from the ca. 300 +/- 200 seconds, which is mostly the
case, and / or the wood moisture content differ considerably from 12%, the
strength values must be modified.

In this paragraph, the influence of the load duration on the


modification factor kmod
this reason Figure 5.5 is extended to figure 6.1.

Introduction to Timber Structures 39


is analyzed. For

Figure 6.1 Effect of the load duration.

Introduction to Timber Structures 40


The failure load curve shown in Figure 6.1 indicates, that a constant result in
load   Fu  F1  Fu
failure at time t1. If the constant load F1    Fu no failure is expected (at any time).
Consequently the permanent load must be lower than   Fu . However, the total load
value, which is the result of 1 1

permanent load and a (number of) variable load, can for a period of time ( t'  t )
exceed the   Fu
without failure of the structure.

At t = 1t' the load value increases from permanent load level to F due to an increased
variable load. 1
The strength is not considerably reduced compared to the short duration
strength (perhaps this strength is slightly reduced due to load history and ageing).
Consequently the load duration should be slightly less than t1 and the failure
load curve is shifted in time.

Occasional increased load levels F1    Fu are therefore no problem and allowed


provided that the cumulative value of the time that F1    Fu does not exceed t1 (
ti  t1 ). The time spans referred to in table 5.3 equal these cumulative values.

The time span ti for load level F1 is determined by the load duration of the variable
load. The variable load is the shortest load in the load combination (permanent +
variable load). Consequently, for the analysis of the structure loaded with permanent
+ variable loading the load duration effects due to the load duration of the variable
loading has to be considered. Since the load duration effects are taken
into account by a modification factor values have to be taken from the variable
kmod , the kmod load.

In general: the modification factor value depends on the shortest load in the
considered load combination.

Generally, load combinations consist out of permanent loads and variable loads
(long, medium or short duration) and often the modification factor associated to the
medium or short hour loading has to be taken into the calculations. Additionally for
construction with a high level of dead load a combination considering only the
permanent load has to be taken into account

On an average level the value   0.56 (see figures 5.5 and 6.1) On the characteristic
value level this
value is higher. Table 6.3 shows a number of values for the 1995-1-1 [9].
modification factor kmod
Introduction to Timber Structures 41
according to EN

Introduction to Timber Structures 42


Table 6.3. Values of kmod .
Material Standard Climate- Load duration class (table 5.3)
class permane long mediu short very
nt m- short
(table
long
5.2)
Sawn EN 14081-1 1 0.60 0.7 0.80 0.90 1.10
timber 2 0.60 0 0.80 0.90 1.10
3 0.50 0.7 0.65 0.70 0.90
0
0.5
5
Glued EN 14080 1 0.60 0.7 0.80 0.90 1.10
laminat 2 0.60 0 0.80 0.90 1.10
ed 3 0.50 0.7 0.65 0.70 0.90
0
wood
0.5
5
LVL EN 14374 , 1 0.60 0.7 0.80 0.90 1.10
EN 14279 2 0.60 0 0.80 0.90 1.10
3 0.60 0.7 0.65 0.70 0.90
0
0.5
5
Plywood EN
636 1 0.60 0.7 0.80 0.90 1.10
Parts 1, 2 and 2 0.60 0 0.80 0.90 1.10
3 3 0.50 0.7 0.65 0.70 0.90
Parts 2 and 3 0
Part 3 0.5
5
OSB EN
300 1 0.30 0.4 0.65 0.85 1.10
1 0.40 5 0.70 0.90 1.10
OSB/
2 0.30 0.5 0.55 0.70 0.90
2 0
OSB/3, 0.4
OSB/4 0
OSB/3,
OSB/4

Factors kh and kl
The factors (height factor) (length factor) are both "volume factors”, described in
kh and kl Chapter 5.
The volume effect is only considered for those material properties showing brittle
failure. Consequently, the volume effect derived from formula (5.5), repeated as
Introduction to Timber Structures 43
formula (6.2), is exclusively for tension both parallel and perpendicular to the grain
and for bending. For bending and tension
parallel to the grain equation (6.2) is reduced to (height factor) (length factor).
the factors kh and kl

0 V 0,2
 max;d
    hV  f [N/mm2] (6.2)
d 0,2 d
 1 
V  ((x, y, z))dV 
 Vh h 

In the denominator of formula (6.2) the volume-integral is elaborated resulting in


one single value for different cases. The values listed in NEN-EN 1995-1-1, 6.4.3
[9], Eurocode 5, are repeated in table 6.4.

Introduction to Timber Structures 44


Table 6.4. Height factors kh and length actor kl .
Material tension parallel to the grain direction and bending
0.2
 150 
Sawn timber with k  700 1.0  kh =  h   1.3
kg/m3
0.1
 600 
1.0  kh =   1.1
Glued laminated timber  h 
LVL (laminated veneer Depending on variations to be determined according to
lumber) EN 14374
Wood-based panels 1.0
Note: the reference length for the length is L = 3000 mm (element length).
factor kl

6.2 Stiffness analysis (Serviceability)


The serviceability of a realised building must also be guaranteed. Movements /
vibrations (for example: floor vibration) must be limited and large deformations must
be avoided. These aspects are related to the so-called Serviceability Limit States
(SLS) for which material properties are given in
e.g. table 6.1.

Generally for calculations in the Serviceability Limit States the average values of
the modulus of elasticity ( Em,0,mean ) and shear modulus ( Gmean ) are used.
Depending on load duration, possibly resulting in creep, and the climate class the
expected deformations are calculated. Figure 6.2 shows the basis for these
calculations.

Figure 6.2 Deformations.

From figure 6.2 it u fin = uinst + (6.3)


follows that ucreep − uc
ep uc
Wit u fin
h unet , fin
uins
t
ucre

Introduction to Timber Structures 45


m] immediate deformation
fina
l t [mm] the creep
def
or h deformation [mm]
mat e the pre camber [mm]
ion
[m final deformation minus pre camber [mm]

The creep deformation is dependent on the load duration. The load duration differs
for different loads (e.g. the load duration for permanent loading is much larger than
for variable loading). From the variable load only small part is permanently present.
This part is taken into account wit a factor  2
(0 ≤ ≤ 1,0), with which the quasi-permanent value of a variable load is calculated,
2 defined in EN
1990 (Eurocode 0) [11]. In principle, the load calculated by multiplying the characteristic
variable load

Introduction to Timber Structures 46


by the factor results in the variable load which is on average permanently present over
2 the
complete life time of the structure. For the permanent tax = 1,0. For variable loads the
follows:  2
values given in table 6.5 [11] apply.
The quasi-permanent value of a variable load indicates the variable load part
resulting in creep deformation. In chapter 5 - Wood properties, part “Wood moisture
content” – it is described, that the wood moisture content (in response to the climatic
conditions) also has an impact on the stiffness and therefore on the deformations. In
practical calculations according to the Eurocodes ([9], [11]) this dependency is only
regarded for the time dependent deformations (creep) by introducing a climate
class dependent creep factor kdef . The can be calculated according to formulas
deformations u fin
(6.4) to (6.6).
u fin,G = uinst ,G  = uinst ,G  for permanent loads G (6.4)
(1+ 2,G  kdef ) (1+ kdef )
u fin,Q = uinst ,Q  (1+ 2,Q  kdef ) for a variable load Q1 (6.5)
1 1 1

u fin,Q = uinst ,Q  ( 0,i + 2,Q  kdef ) for simultaneous variable loads Qi (6.6)
i i i

With  0,i = combination value of simultaneous variable are zero for loading by
loads.  0 and  2
wind, rainwater, temperature and snow so no for these situations no simultaneous
variable loads are present. For variable loads on floors usually only one single
variable load is prescribed. Overall this results normally in only one single
simultaneously variable load in the combination. However for every occurring load
the final deformation needs to be checked. For determining the final deformation ( u
fin )
formula 6.4 to 6.6 can then be simplified in formula 6.7:
u fin = uinst + ucreep − uc = uinst ,G  (1+ kdef )+ uinst ,Q  (1+ 2,Q  kdef )− uc
1 1

u fin = uinst ,G  (1+ kdef )+ uinst ,Q  (keep this in mind!) (6.7)


(1+ 2,Q  kdef )
1 1

Example calculations are made during the exercises. Values (and 1 ) are given in
for  0 and  2
table 6.5 (based on EN 1990 [11]).

Table 6.5.  -factors for variable loads.

Load type Descripti 0 1 2


on
Category A dwellings 0.4 0.5 0.3
Introduction to Timber Structures 47
Category B offices 0.5 0.5 0.3
Category C congresses, meeting places, theatres, 0.6 / 0.7 0.6
conferences 0.4a
Category D shopping 0.4 0.7 0.6
Category E storage 1.0 0.9 0.8
Category H roofs 0.0 0.0 0.0
snow 0.0 0.2 0.0
wind 0.0 0.2 0.0
a for escape routes like stairs: 0.6, other situations: 0.4
Note: the factor 1 is used to determine the so-called frequent value of the
variable loads in case of fire design calculations. The factor 1 is also used to
determine the immediate deformations due to the frequent value of the variable
loads. These deformations are with the load combination

Introduction to Timber Structures 48


according to formula (6.15 b) in EN 1990 [11]. Traditionally there are no
requirements for these deformations in the Netherlands.
Values for for a number of wood products are given in table 6.6, based on EN 1995-
kdef 1-1 [9].

Table 6.6. factors for wood and wood-based materials.


kdef

Climate class
1 2 3
Sawn timber EN 14081-1 0. 0. 2.
6 8 0
Glued laminated EN 14080 0. 0. 2.
wood 6 8 0
LVL EN 14374, EN 0. 0. 2.
14279 6 8 0
Plywood EN
636 0. - -
Part 1 8 1. -
Part 2 0. 0 2.
8 1. 5
Part 3
0. 0
8
OSB EN
OSB/2 2.25 - -
OSB/3, 1.50 2.2 -
OSB/4 5
Note: if it is to be expected, that the wood dries under permanent loading shall
after erection, kdef
be increased with 1,0.

7 Grading: wood quality and strength

7.1 Quality classes and strength classes


Wood is classified into quality classes. For each quality class requirements are set
per wood species related to visual aspects as slope of grain, knots, pith, cracks,
deformations, reaction wood, wane, .... etc. Traditionally, before the wood is traded,
the wood is marked showing the origin (possibly, not always) and from which the
wood quality can be read. Figure 7.1 shows some of these marks.

Introduction to Timber Structures 49


Figure 7.1 Example of marked wood (on the end grain cut of the element).

For wood applications, a distinction must be made between quality classes and
strength classes. For the classification in quality classes, as described above, visual
aspects apply. These visual requirements may also be linked to strength classes.
However, for the classification in quality classes other requirements for the visual
aspects apply than for the classification in strength classes.

Introduction to Timber Structures 50


Quality Classes (not for structural applications: non load-bearing)

The Dutch quality class format is governed by the standards of the KVH 2010 series
(quality requirements for wood), NEN 5461, NEN 5466, .... etc. A distinction is made
in four classes: A t/m D.

Class A: for application with very high demands on the appearance, for example
furniture.
Class B: for applications with high demands on the appearance, for example
constructions with extra demand on the visual aspects.
Class C: common quality, for example timber for regular constructions
Class D: for applications with no requirements on the appearance;
for example, non-load-bearing studs and battens or
products like pallets.

Strength Classes (for structural applications: load bearing)

For calculations according to EN 1995 (EUROCODE – Wood constructions:


Eurocode 5) the so-called characteristic values of the material properties are
necessary. Calculations related to strength and stability (safety) are carried out in the
Ultimate Limit States (ULS) for which the characteristic values are 5% lower values.
Calculations related to deformations are carried out in the Serviceability Limit States
(SLS) for which the characteristic values are the mean values of the modulus of
elasticity and shear modulus.
The characteristic values are taken from a table with strength classes like table 7.1
(extension of table
6.1 – based on EN 338 [7]); the strength class itself is chosen by the structural
designer. The choice is mainly based on availability.
The Dutch strength class format is related to the European and is enshrined in EN
338 [7] for sawn timber and EN 1194 [8] for glued laminated timber.
Table 7.1 shows the strength classes for sawn timber (C-classes with “C” from
Coniferous and
D- classes with “D” from Deciduous). Moreover, the D-classes indicated in table 7.1
are based on the (relatively strong and stiff) tropical deciduous species, also called
hardwoods. Result is, that most deciduous species from the temperate regions
(Poplar - beech - Birch - oak -.... etc.) do not meet the strength valued listed in table
7.1 and are therefore in classified to C-classes.

Introduction to Timber Structures 51


Table 7.1. Strength classes for wood.
Strength class C18 C24 D30 D40 D50 D70 GL24
h
Ultima fm,k N/mm 18 24 30 40 50 70 24
2
te
Limit ft ,0,k N/mm
2 10 14.5 18 24 30 42 19.2
States
ft ,90,k N/mm
2
0.4 0.4 0.6 0.6 0.6 0.6 0.5
(ULS)
fc,0,k N/mm
2 18 21 24 27 30 36 24

fc,90,k N/mm
2 2.2 2.5 5.3 5.5 6.2 12.0 2.5

fv ,k N/mm 3.4 4.0 3.9 4.2 4.5 5.0 3.5


2
k kg/m 320 350 530 550 620 800 385
3
Em,0,k N/mm
2
6,00 7,400 9,200 10,90 11,80 16,80 9,600
0 0 0 0
Seviceabil Em,0,m N/mm 9,00 11,00 11,00 13,00 14,00 20,00 11,50
2
ity Limit ean 0 0 0 0 0 0 0
Em,90, N/mm 300 370 730 870 930 1,330 300
States 2
(SLS) mean
Gmean N/mm 560 690 690 810 880 1,250 650
2

Introduction to Timber Structures 52


• A distinction is made between C-classes ("softwood") and D-classes
("hardwood").
• Any constructive element must be classified in class a strength (no batch
approval allowed based on the approval of random pieces).
• Wood for structural applications can mechanically or visually be graded. If
the wood is visually graded, in the Netherlands this has, for “softwoods” to
be carried out according to NEN 5499 [13]. The class T1 defined in NEN
5499 equals class C defined in the
“KVH”. The class T2 defined in NEN 5499 equals class B defined in the
“KVH”.
• Visually graded Pine, spruce, larch, Douglas (European) and
classified in class T1 according to NEN 5499 [13] meets the
requirements for strength class C18.
• Visually graded Pine, spruce, larch, Douglas (European) and
classified in class T2 according to NEN 5499 meets the
requirements for strength class C24.
• Visually graded Douglas (European) and classified in classes T2
according to NEN 5499: C22
• Oak (Central European), classified in class B accordance to “KVH”: C20
• Meranti (red): strength class D24
• Oak (Polish): D18 / D24 / D30
• Iroko: D24 (unsorted)
• Vitex, Robinia, Sucupira vermelho: D30
• Bilinga: D24 / D50
• Merbau: D30 / D50
• Teak, Iroko (sorted) Sucupira, Itauba, amarelo, Piquia: D40
• Bangkirai, Sapucaia, Angelim vermelho, Denya: D50
• Masseranduba, Cumaru: D60
• Azobé: D70
Note: the strength classes for the different wood species are based on "Wood hand
Strength data [12], a publication of “Centrum Hout” in Almere, the Netherlands.

Introduction to Timber Structures 53


Wood can be assigned to strength classes on the basis of:
1. Visual strength grading
2. Machine strength
grading Both procedures are
non-destructive.

Visual strength grading (structural applications)


Traditionally the strength properties are based on visual aspects. Before assigning
the strength and stiffness values to a piece of wood , the visual characteristics must
be linked to these values. This relationship is established on the basis of
experimental research (by testing in the laboratory). For the most common wood
species used in structural applications, this relationship is established.

The following steps are performed:


• The type of wood is determined
• The Visual characteristics are quantified. In the Netherlands this is for
coniferous wood species carried out according to NEN 5499 (quality
requirements for visually graded coniferous wood species for structural
applications) [13]. This standard provides a classification and associated
requirements on aforementioned visual aspects for strength graded sawn
timber (and wood intended for lamellas to be used in glued laminated wood)
with a rectangular cross-section of the wood species European spruce,
European larch and European pine for load-bearing structures
• On the basis of this classification for solid wood, four strength classes are
distinguished: T0, T1, T2 and T3.
• These four classes are linked to the European strength class system listed in
EN 338 [7] via the European document (European standard) EN 1912 (wood
for structural applications- strength classes-allocation of visual collation
classes and wood species) [14]. Table 7.2 shows this link.

Table 7.2. Link between the strength classes determined according to


NEN 5499 with the European strength classes in EN 338.
NEN 5499 EN 338
T0 C14
T1 C18
T3 C24
T4 C30

Machine strength grading (structural applications)


A parameter which can be measured non-destructively, e.g. mass, modulus of
elasticity, etc. is used for strength prediction. This parameter is called “indicating

Introduction to Timber Structures 54


parameter” (IP). Figure 7.2 shows a relationship between the modulus of elasticity
(IP) and the bending strength.

Introduction to Timber Structures 55


Figure 7.2 Relationship between the modulus of elasticity (IP) and the
bending strength of spruce.

Figure 7.2 is based on bending tests. Figure 7.2 clearly shows that the
relationship between the modulus of elasticity and the bending strength is
ambiguous. Given the measurement of
E = 8000 N/mm2, the bending strength can (on the 5% lower and 95% upper levels)
vary between
fm = 33 N/mm2 and fm = 78 N/mm2 (more than a factor of 2!). However, the
majority of the test results
is close to average making it very unlikely that on the basis of the measured modulus
an extremely low or extremely high bending strength is obtained. Figure 7.2 shows
abundantly clear that on the basis of the indicator (modulus of elasticity) mistakes are
made. The extent to which errors are created can be minimised by considering,
besides the modulus, other indicators: for example: density, visible discolorations,
knots, slope of grain, etc. For capturing these indicators devices are developed which
are useful in the grading process (for example, X-ray measurements of the density,
knot recognition, reaction wood, ... etc.; Laser Scan to capture dimensions, slope of
grain ... wane, etc.).

7.2 Quality marking


Many different indications are stamped on the wood (one example is given in figure
7.1, which is used by the timber traders). Marks indicating the non-structural quality,
e.g. classes A to D according to the “KVH”, as described in part 7.1, are generally
not stamped on the wood.
From safety considerations it is necessary that the correct strength (and stiffness)
values are applied for structurally applied wood. In one way or another the strength
class of any constructive element should therefore be known and consequently
Introduction to Timber Structures 56
every piece of wood intended for structural use must be marked properly.

There are two marks available in the Netherlands:


• KOMO
• CE

Introduction to Timber Structures 57


KOMO
KOMO is the quality mark for the construction market in the Netherlands, originally
founded by the industry to distinguish themselves from other players on the market.
The KOMO certificate is therefore an important tool to improve the quality of supplied
construction elements. The use of construction products/elements/materials supplied
by companies that do not possess the KOMO certificate is therefore greatly reduced.
From Law, in this case the building Act “Bouwbesluit” - part of the “Woningwet”), a
KOMO certificate with certificate is not mandatory. The fact that most construction
products equipped with a KOMO certificate is required is the consequence of the
legal appointments between parties serving as contracts between market parties
(principal – contractor/supplier).
The law, public law, establishes minimum requirements for the realisation of
construction works while a KOMO certificate is a guarantee that these requirements
are generally exceeded.
The requirements to which the product/element/material must meet before the
KOMO certificate is issued are listed in a so-called assessment directive (BRL). All
public safety and health requirements (building Act) are automatically of the BRL.
The industry also imposes additional requirements in order to be able to differentiate
(from your colleague – concurrent). The system for developing and obtaining a
KOMO certificate is in principle shown in Figure 7.3.

Figure 7.3 KOMO.

The system shown in figure 7.3 “guarantees” that an element/product/material


carrying a KOMO certificate automatically meets all requirements from public
law (“Bouwbesluit” requirements).
Introduction to Timber Structures 58
Consequently, an element/product/material carrying a KOMO certificate can be
applied without any additional testing.

Introduction to Timber Structures 59


CE
CE marking is in the (near) future mandatory for all products traded within the
European Union. CE is in contradiction to KOMO mandatory. However, the
underlying documents to CE do not set “limit values” regarding the performance
requirements. The products are developed for a well-defined purpose, e.g. a fire
wall, and the documents describe the properties to be determined and with what
accuracy. The limit values are to be set by the individual Member States of the
European Union.
Consequently it is possible that a product matches the performance requirements
in one Member State and not in another. Construction products carrying CE mark
does therefore not automatically meet the national requirements.
CE marking is based on the so-called construction products regulation in which the
essential requirements (for example, safety, health, and environment) are described.
Taking into account these basic requirements on The European standardisation
Institute (CEN) produces so-called product standards which take these basic
requirements into account. These standards serve as basic documents for CE
marking. For products for which standards are missing (not developed) the
organization EOTA (European Organisation for Technical Approvals) produces so-
called ETAG's (European Technical Approval Guidelines, which serve as a basic
document for an ETA (European Technical Approval). Products for which an ETA
has been derived can be CE marked. The national certification bodies (for example,
SKH = Foundation Inspection Wood - “Stichting Keuringsbureau Hout”) are
members of EOTA.
In CEN standards and ETA’s assessment methods and accuracy requirements for
the relevant properties (e.g. fire resistance). The values obtained are part of the CE
marking and must therefore be enclosed with each delivery. If not, it cannot be
determined whether the properties meet the national requirements and the product
cannot be applied. This is different for the strength and stiffness values of a timber
element intended for structural use since the strength class is printed (stamped) on
these elements as shown in figure 7.4. These stamps can, after thorough research,
even be found on the timber sold in the DIY market. CE marking on elements for
structural use is carried out on the basis of EN 14081 [15].

Introduction to Timber Structures 60


Figure 7.4 CE marking on an element intended for structural use.

Introduction to Timber Structures 61


Figure 7.5 CE.

8 Wood products

Sawn timber and round timber


See powerpoint slides lectures on:
https://canvas.tue.nl/courses/7506 7PPX0 (2018-2) Dimensioning of structures

Glued laminated timber (GlueLam)


See powerpoint slides lectures on:
https://canvas.tue.nl/courses/7506 7PPX0 (2018-2) Dimensioning of structures

Cross laminated timber (CLT)


See powerpoint slides lectures on:
https://canvas.tue.nl/courses/7506 7PPX0 (2018-2) Dimensioning of structures

Laminated Veneer Lumber (LVL or Kerto)


See powerpoint slides lectures on:
https://canvas.tue.nl/courses/7506 7PPX0 (2018-2) Dimensioning of structures

Sheet material
See powerpoint slides lectures on:
Introduction to Timber Structures 62
https://canvas.tue.nl/courses/7506 7PPX0 (2018-2) Dimensioning of structures

Introduction to Timber Structures 63


9 Literature
[1] Josef Kolb. Systems in Timber Engineering. Birkhäuser, Lignum, DGfH,
Switzerland, 2008, ISBN 978-3-7643-8600-4.

[2] Leen Kuiper and Rino Jans (eds). ‘Dutch wood use in image’ ProBos Foundation,
Zeist, 2001.

[3] J. Kuipers. ‘Wood and Wood constructions’ Technische Hogeschool Delft, 1979 (in
Dutch).

[4] Jan f. rijsdijk and Peter b. Laming. ‘Physical and related properties of 145 timbers,
Infomation for practice’ Kluwer Academic Publishers,
Dordrecht/Boston/London, 1994, ISBN 0-7923-2875-2.

[5] Paragraph Foschiand Z.C. Yao. ‘Another look at three duration of load models’
Proceedings of CIB- W18/paper 19-9-1, 1986, Florence, Italy.

[6] Klaassen René. ‘Wood vademecum’ Centrum Hout, Almere, 2018.

[7] EN 338. Wood for structural applications – strength classes. Dutch, Delft Standards
Institute, 2016.

[8] EN14080. Timber structures – Glued laminated timber — Strength classes and
determination of characteristic values. European Committee for Standardisation
(CEN), Brussels, 2013.

[9] EN 1995-1-1. Eurocode 5: design and calculation of wood constructions –


Part1-1: general – common rules and rules for buildings. Dutch
Standardization Institute, 2005.

[10] W. Weibull. ‘A statistical theory of the strength of materials’ 1939.

[11] EN 1990. EUROCODE – basis of the constructive design. Dutch Standardization


Institute, 2002.

[12] Wood wiser ' Strength ', a publication of the Data Centre Wood to
Almere, located under
http://www.houtinfo.nl/pdf/Houtwijzer%20Sterktegegevens%20van%2
0hout.pdf

[13] NEN 5499. Requirements for visually graded softwood for constructive
applications (in Dutch). Dutch Standardisation Institute, Delft, The
Netherlands, 2007.

[14] EN 1912. Structural timber – Strength classes – Assignment of visual


Introduction to Timber Structures 64
grades and species. European Committee for Standardisation (CEN),
Brussels, 2012.

[15] EN 14081. Timber structures - Strength graded structural timber with


rectangular cross section. European Committee for Standardisation (CEN),
Brussels, 2016.

Introduction to Timber Structures 65


Part 2: DESIGN OF WOODEN LOAD-BEARING STRUCTURES

10 Introduction
When designing you start with a course draft, which has to be refined. Designing is
a multidisciplinary process. Input from every relevant discipline is necessary. Within
the building industry all kind of construction works within the built environment,
usually buildings, and civil engineering objects.

This part of “Introduction to Timber Structures” focuses on buildings in which


architecture, urban planning, building physics, accommodation of installations and
structural design are involved in the multidisciplinary designing process.

The design process needs a systematic approach. The first design at the start of the
building process is extremely important. In this phase, the preliminary design phase,
the design effort pays off. Further on in the design process the effect of efforts
reduces. This is illustrated in figure 10.1.

Figure 10.1 Influence on the design in the various phases of the


construction process.

During the design phase a (large) number of possibilities should quickly and
efficiently be assessed. Experienced designers use their experience where rules of
thumb and/or simplifications (for example, the reduction of complex structures into
easy to understand static structures) play an important role. The rules of thumb are
based on earlier experience. In this part of the course rules of thumb are presented
for dimensions based on [1]. Only profiles with a rectangular cross-section are
considered.

Generally it can be stated that:


• Bending moments require relatively more material (larger cross section
Introduction to Timber Structures 66
dimensions) than axial forces. In an economical design the bending, moments
are reduces as much as possible. This means, that the geometry, the position
of the neutral axis, follows as close as possible the so- called “line of thrust”.
An example of this is the (parabolic) arch structure shown in figure 10.2,
where the bending moments due to an evenly distributed load are zero (no
bending moments).

Introduction to Timber Structures 67


Figure 10.2 Parabolic truss (follows almost the pressure line for
evenly distributed layer.

On the basis of figure 10.2 three hinge arches can be optimised. Since hardly
ever an element loaded in bending is subject to constant moment loading,
material is saved by introducing non- prismatic elements. In these cases an
option can be to vary the cross section dimensions in relation to the variation
in bending moment. Figure 10.3 shows an example.

Figure 10.3 At the custom profile height gradient moments.

Non prismatic elements must be designed with great care since stresses
perpendicular to the grain are introduced and wood is significantly weaker
perpendicular to the grain than parallel to the grain. Especially cases where
tension stresses perpendicular to the grain are introduced should be avoided
(to avoid brittle failure).
• High narrow cross sections are usually more economical than low wide
cross sections. It should be noted, that this is especially true for glued
laminated bending elements, for which
h
cross sections  8 are common. Sawn timber profiles are generally stocky
with (from
b
b x h = 59 x 146 to 69 x 269). For rafters in prefabricated roof elements

Introduction to Timber Structures 68


sawn timber cross sections are more slender: approximately 30 ≤ b ≤ 40
mm with height ≤ 286 mm).
• Since higher strength and stiffness properties result in smaller cross section
dimensions, it can be efficient to find an optimal match between these
properties and the application conditions.
• For elements in bending, the most outside fibres are stressed most.
Therefore, for glued laminated timber so-called combined cross sections are
produced (so-called GLc – in which “GL” denotes Glued Laminated and “c”
denotes combined) for which the outer lamellas are chosen from a higher
strength class.

Introduction to Timber Structures 69


• From a durability point of view (for prevention degradation due to fungi)
projects should be carried out, so that exposure to wind and weather is
minimised. For example, for housing projects with a number of homes in a
row, this means vertical realisation (covered by the roof structure as soon as
possible) instead of the usual horizontal construction, which is the usual way
in the realisation of brick houses.
• To design a building, we are dealing with a main load carrying structure, a
secondary load carrying structure, foundation, etc. Structures to ensure
stability are also needed. The cost of the load carrying structure is a sum of
the costs of these elements. Usually the in between distance of the main load
carrying elements is about 5 meters. For larger distances the secondary
elements should be glued laminated, which is considerably more expensive
than a secondary support structure of sawn timber. However, the cost of the
main support system reduces if the distance is increased. For this reason, it is
advisable to choose in between distances in the range outside 5 meters to 7
meters.

11 Rules of thumb for girders and columns

11.1 Rules of thumb for girders


The main characteristic of structural components in sawn timber (also called solid
timber) is the limited dimensions in terms of both cross section dimensions and
length. Usual trade dimensions of sawn timber elements show a length range from
1.80 meters in ascending steps of 30 cm up to a maximum of 6,0 meters. However,
as mentioned in Chapter 1, it is not expedient to design lengths greater than
5.0 meters because of limited availability resulting in higher costs. In any case it is
advisable to inform on the availability in the market if application of elements with
length values > 5 meter is intended. Extension of the length above 5.0 meters can be
achieved by full cross section finger jointing, which is regularly applied for multi
supported beam elements (and purlins) and rafters in roof structures.

The application of sawn timber is limited to relatively small spans in roofs and
floors, storey high columns and facade poles.

Glued Laminated timber on the other hand, is much wider applicable. The
dimensions are not limited by the tree dimensions.

For standard timber floors and roofs a first estimation of the cross section
dimensions can be obtained using rules of thumb as indicated in figure 11.1.

Introduction to Timber Structures 70


Figure 11.1 Rules of thumb for girders (beams in bending).

Note: The cross section width for the sawn timber beam types given in figure 11.1can
be calculated
c.t.c.
according to b = for floor beams and b =
10 c.t.c. for roof beams. (with c.t.c.. = beam
15
h h
b = to .
spacing). The cross section width for glued laminated 8 5
elements varies from
In many cases deformation limits are governing the design and consequently the
cross sectional dimensions can be corrected by keeping b  h3 constant. In those
cases that strength is governing, b  h2 must be kept constant.

Introduction to Timber Structures 71


Example: flat roof with beam span L = 4.8 m and beam spacing c.t.c.. = 0.6 m

4800
Design bar dimensions h= = 250 mm
19
600
b= = 40 mm
15

Introduction to Timber Structures 72


40  2503
Cross section dimensions (standard available): h  3 = 208 mm : 69 x 219 mm
69
Of course, after some simplification, the cross section dimensions can be determined
by a global
calculation. Girders (or beams) are loaded in bending. A design on strength is carried
out in the so- called Ultimate Limit State (ULS) and the cross section height can be
determined according to formula (2.1).
6M
h  b  f d mm (11.1)
m,0,d

With b width [mm]


h height [mm]
Md design value of the bending moment [Nmm]
fm,0,d design bending strength according to equation (6.1). In the
Netherlands the strength and stiffness properties for glued laminated
timber are usually taken from the GL24h
24
1.2  0, 9 = 17.3 N/mm . For sawn timber the strength
strength fm,0, 2
class: d = 5 values are
18
taken from the C18 or C24 strength f =  0, 9 = 12.5 N/mm2 or
classes:
m,0,d 1.3
24 2
 0,9 = 16, 6 N/mm
respectively.
1,
3

Fore uniformly loaded beams on two or three supports the governing design bending
moments M d
are given in figure 11.2.

Figure 11.2 Values for M d


Introduction to Timber Structures 73
(design of girders on the basis of strength; ULS).

In many cases, certainly for most beams on more than two supports, deformations
are governing the design. Deformations are assessed in the so-called Serviceability
Limit States (SLS). The required cross section dimensions can be calculated
according to formula 11.2.
1
I   b h3 = mm4
C  qd an h  3 12  [mm] (11.2)
d C  qd
12 E b E

Introduction to Timber Structures 74


With C coefficient [mm3]. For girders on two and three support points
loaded uniformly distributed with qd , C is given in figure 11.3.
E modulus of elasticity (average value E0,mean ) [N/mm2]. For glued laminated
timber the
stiffness values are taken from strength class GL24h: N/mm2. For
E0,mean = 11, 500
sawn timber the strength and stiffness values are taken from
strength class C18 or C24: E0,mean = 9, 000 N/mm2 respectively.
E0,mean = 11, 000 N/mm2.

Figure 11.3 Girder (beam) designs based on stiffness (SLS).

For a preliminary design phase the loads can be taken from figure 11.4.

ULS: qd =  G  Gk +  Q  Qk
 G = 1.08 ,  G = 1.35 (consequence class CC1)
SLS: qkr = q fin = Gk  (1+ kdef )+ Qk  (1+ 2  kdef )
kdef = 0.6 (climate class 1)
floor Roof
s s
dwellings office buildings slope fla
d t
 2 = 0.3  2 = 0.3  2 = 0.0  2 = 0.0
Gk = 0.50 kN/m2 Gk = 0.50 kN/m2 Gk = 0.75 kN/m2 Gk = 0.50 kN/m2
(tiles)
Introduction to Timber Structures 75
Qk = 1.75 kN/m2 Qk = 2.50 kN/m2 Qk = 0.80 kN/m2 Qk = 1.00 kN/m2

Figure 11.4 Loads.

Introduction to Timber Structures 76


Example: flat roof with beam span L = 4.8 m and beam spacing c.t.c.. = 0,6 m
 2 = 0.0 , kdef = 0.6 , strength class C24
qd = c.t.c. ( G  Gk +  Q  Qk ) = 0.6  (1.08  0.50 +1.35 1.0)=1.13 kN/m
1 1
M =  q  L2 = 1,13 4,82 = 3.3 kNm
d 8 d 8
ULS
6 M 6 3.3106
b = 44 mm : h  d = = 165 mm
b  fm,0,d 44 16.6

qkr = c.t.c. (Gk  (1+ kdef )+ Qk  (1+ 2  kdef )) = 0.6 (0.50 (1+ 0.6)+1.00 (1+
0.0 0.6)) = 1.08 kN/m
4,800
 = = 0, 004  4,800 = 19.2 mm
SLS max 250
5 L4 q kr 5 4,8004 1.08 9 · mm4
I   =   = 0.035310
384 max E 384 19.2 11, 000
12  C  q 12  0.0353109
b = 44 mm : h  3 kr = 3 = 213 mm
bE 44

The calculation in the Serviceability Limit States results in a higher height value.
Consequently SLS is governing. Cross section dimensions (standard available): 44 x
219 mm.

Introduction to Timber Structures 77


11.2 Rules of thumb for columns
Columns are loaded in compression. In many application a horizontal load result in
additional bending (e.g. facade post). The rules of thumb for elements loaded in
compression and bending are given in figure 11.5.

Buckling
L
h=
b  h = 1+ 3 
20 F e
r
 

5 h
sawn timbe
ated timber b  h = F 1+ 3 e 
 

7 h
glued lamin
constant
2
with b  h =
Figure 11.5 Rules of thumb for elements loaded in compression (and
bending) [1].

Example: flat roof with beam span L = 4.8 m and beam spacing c.t.c.. = 0.6 m,
strength class C24
( )
qd = c.t.c.  G  Gk +  Q  Qk = 0.6  (1.08  0.50 +1.35 1.0)=1.13 kN/m

1 1
Md =  qd  L2 = 1,13 4,82 = 3.3 kNm
8 8
Additional the element is subjected to a design compression load of Fd = 100 kN. The calculation is
carried out in the Ultimate Limit State (ULS) only.
4,800
h= = 240 mm
20
M 3.3106
e= = = 33 mm
F 100103
F  e  100 103  33 
b=  1+ 3  =  1+ 3 = 118 mm
h5  h 240 5  240 

The profile 118 x 240 mm is not a commercial size. With b = 121 mm it follows that

118  2402
h = 237 mm. Commercial size: 121 x 245 mm.
121
Introduction to Timber Structures 78
12 Rules of thumb for three-hinge-frames
The in between distance, truss shape (curved/fragmented), the span, the gutter
height, and the ridge height and slope, determine the dimensions of the frame. This
is graphically shown in figure 12.1, in which figure 10.2 (“line of thrust”) plays a key
role.

Figure 12.1 Three-hinge kinked frame (left half) and curved frame (right
half).

The dimensions of the cross section of three hinge frames are mainly
determined by bending moments. The axial forces are of minor importance.

Figure 12.1 also shows the so-called “line of thrust” for uniform distributed loading. If
the system line of the frame coincides with this “line of thrust” no bending moments
develop; the load is transferred by the frame by axial forces only indicated with "N" in
Figure 12.1. This is the roughly case for a parabolic arch truss.
If the truss system line does not coincide with the “line of thrust” bending moments
develop which can be calculated by multiplying the axial load by the distance
between the system line and the “line of thrust”; e.g. the bending moment in the
cranked corner (figure 3.1 – left) equals M = ek  N . The deviations are indicated in
figure 12.1 with " ek " for the cranked truss (left half) and " eb " for the curved truss
(right half). Since " ek  eb " the bending moments developed in the frame with the
cranked corner (left) are larger than in the frame with the curved corner (right).
Introduction to Timber Structures 79
Consequently, frames with cranked corner require more material than frames with a
curved corner. This is reflected in the application of the rules of thumb given in figure
12.2.

The cross section dimensions obtained by the rules of thumb in figure 3.2 can be
transformed into available dimensions by keeping b h2 = constant.

Introduction to Timber Structures 80


Figure 12.2 Rules of thumb for three-hinge frames.

Introduction to Timber Structures 81


13 Structural detailing
Structural detailing is the analysis of all forces on the joint to facilitate these forces to
pass the joint. The most efficient way to achieve this is by contact pressure. In those
cases connectors are only necessary to position the elements and not for load
transfer. However, that is not always possible; for example, the load transfer in truss
joints have to be taken care for by mechanical fasteners (or glued connections) for
which mostly dowel type fasteners are used. Figure 13.1 shows a wide variety of
dowel type fasteners available in the market nowadays.

Figure 13.1 Dowel type fasteners.

Introduction to Timber Structures 82


In the preliminary design phase the strength of a connection with dowel type
fasteners can be estimated by the equations shown in figures 13.2. and 13.3.

Figure 13.2 Load carrying capacity of a connection with a single dowel


type fastener, depending on the timber thickness (t1 and t2)

Figure 13.3 Rules of thumb for determining the design load


carrying capacity of connections with dowel type
fasteners.
Note: n = number of fasteners (parallel or perpendicular to the
Introduction to Timber Structures 83
grain)

Introduction to Timber Structures 84


Figure 13.3 (continued) Rules of thumb for determining the design
load carrying capacity of connections with dowel type
fasteners.
Note: n = number of fasteners (parallel or perpendicular to the
grain)

Example: calculation of the maximum normal force N for the connection below

Step 1: Minimum dimensions:


h  23d + (m −1)  4d = 2312 + (4 −1)  412 = 216 mm
Introduction to Timber Structures 85
For table A the following condition should apply: t1  5d = 512 = 60 mm → t1 = 60 mm OK
For table A the following condition should apply: t2  4.5d = 4.512 = 54 mm→ t2 = 80 mm OK
L  2(2 7d + (n −1)  7d ) = (n +1) 14d = (n +1) 14d = (3+1) 1412 = 672 mm

Step 2: Determine design force Fv, Rd


Graph A in figure 13.3 Fd = n ef  40d 2 n  2: nef = 0,75

Fv,Rd = nef  40d 2 = 0.75 40122 = 4,320 N per shear plane

Fv,Rd ,tot = nshear plane  nbolts  Fd = 212 4,320 = 103.68103 N = 103.7 kN

Step 3: Unity Check of the forces in the connection


N Ed 100 103
UC= = = 0.96  1.00 OK
Fv,Rd ,tot 103.7 103
Note: check the normal stresses in the timber elements, reduced by the holes for the bolts, as an
additional assignment

Connections in tension are, in the past, also realised with so-called carpentry joints
(traditional timber connections). An example is shown in figure 13.4 (Angera castle,
Lago Majore, Italy).

Figure 13.4 Tensile connection using a so called carpentry connection.

The effectiveness of the connection shown in figure 13.4 is low because only a
small portion of the timber element is activated for load transfer (only the
compressed area). The effectiveness of the connection shown in the example on
page 50 is much bigger. This is the main reason the so-called carpentry
connections are not applied often nowadays. However, aesthetic reasons and
because of the fact that due to automatic production processes the prefabrication
is much faster and accurate than they used to be, carpentry connections revived
up to a certain extend.

Introduction to Timber Structures 86


Examples where the forces are mainly transferred by contact pressure, the most
efficient way of detailing, are developed for prefabricated roof structures. One
possible detail is shown in figure 13.5.

Introduction to Timber Structures 87


Figure 13.5 One of the details developed for prefabricated roof structures.

The axial and shear loads, indicated in Figure 13.5, are transferred into vertical and
horizontal loads H and V introducing compression and tension perpendicular to the
fibre and rolling shear in the wall plate. Additionally, bending stresses are introduced
into the F-shaped steel element. The screw only serves to position all elements and
for transferring an upward shear load due to wind suction (loading the screw in
withdrawal).

Figure 13.6 shows an example of a post and beam structure where the mechanical
fasteners (in this case: dowel type fasteners) play a key role in the load transfer.

Introduction to Timber Structures 88


Figure 13.6 Post-and-beam connection (church in Daarle, Overijssel, The
Netherlands).

In principle the beam shear force results in a compression force in the column. The
connection has to transfer this force. The shear force is transferred to the T-shaped
steel element by fasteners “A” from which the force has to be transferred to the
column central axis. Obviously an eccentricity, resulting in a bending moment “M”,
develops. The bending moment due to the eccentricity in the connection (in the case
shown in figure 13.6 M = 140 · shear force) is transferred by fasteners “B”, through
which de beam is loaded parallel to the grain.
Since the fasteners “B” transfer load parallel to the fibre direction, the holes for these
bolts in the T- shaped steel plate can be oval shaped (with the large oval axis
vertically) allowing the timber to shrink and swell without developing tension stresses
perpendicular to the grain. This becomes important when the distance between the
bolts “B” in glued laminated timber exceeds 500 to 600 mm and in sawn timber 180
to 200 mm (the reason for this difference between glued laminated timber and sawn
timber is that during erection generally the wood moisture content of sawn timber is
much higher).
Obviously the bending moment on the column resulting in bending stresses is shown
in figure 13.6 as M = eccentricity · shear force.

Introduction to Timber Structures 89


14 Stability
In general, timber construction elements like beams and columns are designed to
transfer load in one plane. To ensure that the load can actually be transferred, the
stability of these elements must be assured. Stability elements are often combined
with structures for horizontal wind load transfer.

Walls, roofs and floors are loaded perpendicular to the plane. Forces due to stability
load these elements in plane for which they can be designed properly (by activating
the sheet material with which these elements are usually finished). For walls and
roofs, however, often special bracing elements are added.

For roofs of houses the sheet material is, however, mostly activated. This is illustrated
in figure 14.1.

Figure 14.1 Stability provided by the roof sheet material.


Note: the sheet material is loaded due to the wind load on the
gable wall.

Legend to figure 14.1: N1 Normal force due to the wind load perpendicular to
the Gables [kN].
qw = wind load in kN/m2.
A force parallel to the Gables needed for moment equilibrium
[kN]

The connection between the gable and roof sheet material must be designed on
a force which is a combination of the forces “N1” and “A”. The connection
between the individual elements in the roof must be able to transfer the shear
force A and a portion of the load N1.

Bracing systems can be carried out using inclined steel bars, exclusively on
loaded in tension (e.g. the cross bracing system shown in figure 14.4). They can
also be realised with inclined timber elements (also cross bracing) of which half of
Introduction to Timber Structures 90
the elements is loaded in tension and the other half in compression; see figure
14.3.

No eccentricities should be introduced. This is illustrated in Figure 14.2: it is useful


to situate the steel rods underneath the beams. However, than an eccentricity is
introduced resulting in an eccentricity moment causing the main girder to rotate:
figure 14.2 (a). Figure 14.2 (b) shows an example where this eccentricity, and its
negative effects, is overcome.

Introduction to Timber Structures 91


Figure 14.2 Connection of the (steel) cross bracing system.

Wooden bracing, whether or not in cross bracing, have the advantage that the
elements are able to transfer both tension or compression resulting in half of the
force in the elements (compared to the situation above, where, due to the steel
cross bracing, the elements can be loaded in tension only). Consequently, the
connections can be designed on half of the force as well. Figure 14.3 shows a
cross bracing with timber elements.

Figure 14.3 Cross bracing with timber elements (detail).

Because of the element cross sectional dimensions they need to be notched at the
intersection (in the middle). This notch is a disadvantage regarding the buckling
behaviour of the element in compression (which is, however, supported by the
element in tension through the bolt indicated in figure 14.3).
The notches cause an eccentricity, resulting in a bending moment, resulting in
deformation of both the element in tension and compression. Consequently, the
deformation of the element in compression reduces the buckling resistance. The
annoying thing is that both the elements in tension and in compression tend to
deform in the same direction.

When assessing the stability in plane of the cross bracing system, half of the
Introduction to Timber Structures 92
element length (see figure 14.3) can safely be taken as buckling length. In the other
direction, perpendicular to the cross bracing plane, this somewhat more nuanced; in
this direction 0.75 times the element length can safely be taken as buckling length.
Due to this larger buckling length, the orientation of the elements with the

Introduction to Timber Structures 93


largest cross sectional dimension (height) perpendicular to the cross bracing plan is
the most effective orientation.

For halls (many main elements parallel to each other, the best location for the
cross bracing is immediately after the gables. In that case the wind load on the
gables is directly transferred into the roof bracing system. However, gable frames
are often not applied. In that case the cross bracing is shifted away from the gable.

In order to transfer the gable wind load directly into the roof bracing system it is
advisable to match the purlins in the roof structure with the façade poles.

The "width" of the bracing, in figure 14.4 indicated with "L", must be large
enough to prevent overloading of the frames / trusses.

Figure 14.4 Cross bracing (wind loading).

In the bracing section of the building, as shown in figure 14.4, vertical loads develop
for slopes > zero. These loads, in figure 14.4 indicated with V, increase with
increasing slope; for a flat roof the effect is not existing. The effect reduces with
increasing “L” (see figure 14.4).
Table 14.1 gives a number of recommendations (rules of thumb) to keep this effect
acceptable.

Table 14.1. Minimum “L” (figure 14.4).


 = slope L

  25o L 10 m
25    L 15 m
35o
  35o L  20 m

Note:  L total “width” of the cross bracing; e.g.  L in figure 5.5 is the total “width”

Introduction to Timber Structures 94


of the three bracing systems. The cross bracings are spread over the roof
structure. Choose a maximum spacing between the cross bracings of about 20 to
25 meters.

Introduction to Timber Structures 95


Figure 14.5 Location of the cross bracings.

Cross bracing systems transfer wind loads and support elements which are loaded
in compression and/or in bending. For elements in bending cross bracings are most
effective when supporting the cross section zone subjected to compression.
However, timber is a very light weight material also applied in large span flat roof
structures where wind suction is higher than the dead weight of the structure.
Consequently the compression zone is not located near the roof, where it is rather
easy to support, but at the opposite side causing this zone to buckle (this
phenomenon is called torsion buckling). To avoid torsion buckling the cross
section has to be supported against rotation. Figure
14.6 shows some possibilities.

Figure 14.6 Support of the cross section against rotation (avoiding torsion
Introduction to Timber Structures 96
buckling).

Introduction to Timber Structures 97


15 Literature

[1] W.J. Raven. ‘Rules of thumb for determining of floors, beams and columns in
wood, steel and concrete' Faculty of civil engineering of the Technical
University in Delft, 2003.

[2] EN 1995-1-1. Eurocode 5: design and calculation of wood constructions –


part 1-1: general – common rules and rules for buildings. Dutch
Standardization Institute, Delft, 2007.

Introduction to Timber Structures 98


Annex 1: Calculation example: Unity Checks of the stresses in a single span
beam

As a calculation example a floor beam on two supports and a span of 10 m will be


checked. No normal forces are taken into account. The loads and safety factors
(Consequence Class 1) are:
- Gk = 4.0 kN/m1 (including dead load of the beam),  g = 1.2

- Qk = 8.0 kN/m1 (including dead load of the beam),  q = 1.35

For the assessment of the wooden beam material and modification factors are
needed to determine the design value for strength:
- Partial factor (  M ) for the material properties for glulam:  M = 1, 25
- Modification factor ( kmod ) for climate class and the duration of the load:
kmod = 0,9 (climate class 1, short term action)
- Modification factor ( kdef ) for creep: kdef = 0.6 (climate class 1, glued
laminated timber)
- Factor for variable load:  2 = 0.3 (Category A: dwellings)

Glulam beam with dimensions: width x height = 130 x 680 mm (h: L/16, w: h/5)
Strength class GL28h For indoor conditions the thickness of the lamellas is 40 mm,
for outdoor conditions the thickness is 27 mm. A total of 17 lamellas gives a height of
680 mm.

Bending strength (  m,d  fm,d )


The design value of the load is:
qd =  g  Gk +  q  Qk = 1.2  4.0 kN/m1
+1.358.0 = 15.6

The design value of the bending moment is:


1 1
M =  q  L2 = 15.6 102 = 195 Nmm
kNm = 195106
d 8 d 8
Introduction to Timber Structures 99
The moment of resistance:
1 1
W =  w  h2 = 130  6802 = 10, 019 103 mm3
6 6

The design value of the bending stress is calculated as:

Introduction to Timber Structures 100


Md 195 106
 = = = 19.5 N/mm2
m,d 10, 019 103
W

The design value of the strength is:


28.0
fm,d m,k mod =  0.90 1.0 = 20.2 N/mm2
f
1.25
=  k  kh
M

Check of the stresses in the Ultimate Limit State (ULS):


m,d 19.5
Unity = ≤ 1.00 OK
Check: = 0.97
fm,d 20.2

The beam satisfies the conditions with a marge of 3%.

Shear stress (  v,d  fv,d )


The theory of the calculation of the shear force and the associated shear stresses is
also based on a completely elastic behavior of the cross-section. Thus, the shear
stress curve is assumed to be
parabolic about the height. For the maximum shear in the heart of the beam,
stress  v,d therefore,
applies:
3 V
 =  d = fv,k  k
f 
v,d
2bh v,d mod
M

For the calculation example the design shear force is:


1 1
V =  (  G +   Q ) L =  (1.2 4.0 +1.358.0)10 = 78.0 kN
d 2 g k q k 2

For the shear stresses it follows:


3 V 3 78103
 =  d =  = 1.32 N/mm2
v,d 2 b  h 2 130  680

The design value of the allowable shear stress is:


fv,k 3.5
f = k =  0.9 = 2.52 N/mm2
v,d  mod
1.25
M

Introduction to Timber Structures 101


v,d 1.32
Unity = = ≤ 1.00 OK
Check:
0.52
fv,d 2.52

Introduction to Timber Structures 102


Deflection ( u fin  wlimit )
→ q fin = Gk  (1+ kdef )+ Qk  (1+ 2  kdef )
u fin = uinst ,G  (1+ kdef )+ uinst ,Q 
(1+ 2,Q  kdef )
1 1

q fin = Gk  (1+ kdef )+ Qk  (1+ 2  kdef )= 4.0  (1+ 0.6)+ 8.0  (1+
0.3 0.6) = 15.84 kN/m1. This is the load

including all the creep factors! No adjustments are now needed anymore for the
stiffness (elastic modulus E0,mean ).

The moment of inertia of the gluelam beam is:


1 1
I =  b  h3 = 130  6803 =
· mm4
340, 634 104
12 12

The final deflection, including creep, now becomes:


u =
5q  5 15.84 10, 0004 48.1 mm
=  =
 fi L
n 4
fin
38 E  384 12, 600  340, 634 104
4 I

wlimit = 0.004 L = 0.00410, 000 = 40.0 mm

u = 48.1 =
Unity ≥ 1.00 not OK
fin 1.20
Check:
wli 40.1
mit

Adjustments are needed to increase the stiffness to satisfy the conditions for the
deflection.

Introduction to Timber Structures 103


Annex 2: Tables Eurocode 5

Table 2.1. Material factors  m .


materi Material factor
al m
sawn timber 1.30
glued laminated 1.25
timber
LVL, plywood, OSB 1.20
connections 1.30
metal plate 1.25
connectors

Table 2.2. Height factors kh and length actor kl .


Material tension parallel to the grain direction and bending
0.2
 150 
Sawn timber with k  700 1.0  kh =  h   1.3
kg/m3
0.1
 600 
1.0  kh =   1.1
Glued laminated timber  h 
LVL (laminated veneer Depending on variations to be determined according to
lumber) EN 14374
Wood-based panels 1.0
Note: the reference length for the length is L = 3000 mm (element length).
factor kl

Table 2.3. Climate classes.


Climate ave Description
class rage
[%]
1 12 Standard indoor conditions
2 20 Outdoor, covered structures
3 >20 - poorly ventilated spaces (indoor)
- fully exposed to outdoor conditions (not
covered)
- structures in and underneath water

Table 2.4. Load duration classes.


Load duration class Cumulative duration of the Examples
characteristic load
Permanent Longer than 10 years Dead load
Long 6 months - 10 years Storage

Introduction to Timber Structures 104


Medium-Long 1 week - 6 months Life loads on floors
Short Less than 1 week Snow, wind (The
Netherlands)
Instantaneous Accidental load, wind
(Belgium)

Introduction to Timber Structures 105


Table 2.5. Values of kmod .
Material Standard Climate- Load duration class
class permane long mediu short very
nt m- short
(table
long
5.2)
Sawn EN 14081-1 1 0.60 0.7 0.80 0.90 1.10
timber 2 0.60 0 0.80 0.90 1.10
3 0.50 0.7 0.65 0.70 0.90
0
0.5
5
Glued EN 14080 1 0.60 0.7 0.80 0.90 1.10
laminat 2 0.60 0 0.80 0.90 1.10
ed 3 0.50 0.7 0.65 0.70 0.90
0
wood
0.5
5
LVL EN 14374 , 1 0.60 0.7 0.80 0.90 1.10
EN 14279 2 0.60 0 0.80 0.90 1.10
3 0.60 0.7 0.65 0.70 0.90
0
0.5
5
Plywood EN
636 1 0.60 0.7 0.80 0.90 1.10
Parts 1, 2 and 2 0.60 0 0.80 0.90 1.10
3 3 0.50 0.7 0.65 0.70 0.90
Parts 2 and 3 0
Part 3 0.5
5
OSB EN
300 1 0.30 0.4 0.65 0.85 1.10
1 0.40 5 0.70 0.90 1.10
OSB/
2 0.30 0.5 0.55 0.70 0.90
2 0
OSB/3, 0.4
OSB/4 0
OSB/3,
OSB/4

Table 2.6.  -factors for variable loads.


Load type Descripti 0 1 2
on
Category A dwellings 0.4 0.5 0.3
Category B offices 0.5 0.5 0.3
Introduction to Timber Structures 106
Category C congresses, meeting places, theatres, 0.6 / 0.7 0.6
0.4 a
conferences
Category D shopping 0.4 0.7 0.6
Category E storage 1.0 0.9 0.8
Category H roofs 0.0 0.0 0.0
snow 0.0 0.2 0.0
wind 0.0 0.2 0.0
a for escape routes like stairs: 0.6, other situations: 0.4
Note: the factor 1 is used to determine the so-called frequent value of the
variable loads in case of fire design calculations. The factor 1 is also used to
determine the immediate deformations due to the frequent value of the variable
loads. These deformations are with the load combination according to formula
(6.15 b) in EN 1990 [11]. Traditionally there are no requirements for these
deformations in the Netherlands.

Introduction to Timber Structures 107


Table 2.7. factors for wood and wood-based materials.
kdef

Climate class
1 2 3
Sawn timber EN 14081-1 0. 0. 2.
6 8 0
Glued laminated EN 14080 0. 0. 2.
timber 6 8 0
LVL EN 14374, EN 0. 0. 2.
14279 6 8 0
Plywood EN
636 0. - -
Part 1 8 1. -
Part 2 0. 0 2.
8 1. 5
Part 3
0. 0
8
OSB EN
OSB/2 2.25 - -
OSB/3, 1.50 2.2 -
OSB/4 5
Note: if it is to be expected, that the wood dries under permanent loading shall
after erection, kdef
be increased with 1,0.

Introduction to Timber Structures 108


Table 2.8. Strength classes for wood.
Strength class C18 C24 D30 D40 D50 D70 GL24
h
Ultima fm,k N/mm 18 24 30 40 50 70 24
2
te
Limit ft ,0,k N/mm
2 10 14.5 18 24 30 42 19.2
States
ft ,90,k N/mm
2
0.4 0.4 0.6 0.6 0.6 0.6 0.5
(ULS)
fc,0,k N/mm
2 18 21 24 27 30 36 24

fc,90,k N/mm
2 2.2 2.5 5.3 5.5 6.2 12.0 2.5

fv ,k N/mm 3.4 4.0 3.9 4.2 4.5 5.0 3.5


2
k kg/m 320 350 530 550 620 800 385
3
Em,0,k N/mm
2
6,00 7,400 9,200 10,90 11,80 16,80 9,600
0 0 0 0
Seviceabil Em,0,m N/mm 9,00 11,00 11,00 13,00 14,00 20,00 11,50
2
ity Limit ean 0 0 0 0 0 0 0
Em,90, N/mm 300 370 730 870 930 1,330 300
States 2
(SLS) mean
Gmean N/mm 560 690 690 810 880 1,250 650
2

Introduction to Timber Structures 109


• A distinction is made between C-classes ("softwood") and D-classes
("hardwood").
• Any constructive element must be classified in class a strength (no batch
approval allowed based on the approval of random pieces).
• Wood for structural applications can mechanically or visually be graded. If
the wood is visually graded, in the Netherlands this has, for “softwoods” to
be carried out according to NEN 5499 [13]. The class T1 defined in NEN
5499 equals class C defined in the
“KVH”. The class T2 defined in NEN 5499 equals class B defined in the
“KVH”.
• Visually graded Pine, spruce, larch, Douglas (European) and
classified in class T1 according to NEN 5499 [13] meets the
requirements for strength class C18.
• Visually graded Pine, spruce, larch, Douglas (European) and
classified in class T2 according to NEN 5499 meets the
requirements for strength class C24.
• Visually graded Douglas (European) and classified in classes T2
according to NEN 5499: C22
• Oak (Central European), classified in class B accordance to “KVH”: C20
• Meranti (red): strength class D24
• Oak (Polish): D18 / D24 / D30
• Iroko: D24 (unsorted)
• Vitex, Robinia, Sucupira vermelho: D30
• Bilinga: D24 / D50
• Merbau: D30 / D50
• Teak, Iroko (sorted) Sucupira, Itauba, amarelo, Piquia: D40
• Bangkirai, Sapucaia, Angelim vermelho, Denya: D50
• Masseranduba, Cumaru: D60
• Azobé: D70
Note: the strength classes for the different wood species are based on "Wood hand
Strength data [12], a publication of “Centrum Hout” in Almere, the Netherlands.

Introduction to Timber Structures 110


Annex 3: Floor beams: ratio span versus beam height (rules of thumb)

Gk = 0, 46 kN/m2
Qk = 1.75 + 0.50 = 2.25

kN/m2 Center to center: 600


mm
L
u = 0, 004 
fi L =
n 250
q fin = Gk  (1+ kdef )+ Qk  (1+ kdef  2 )= (0.46  (1+ 0.6)+ 2, 25 (1+ 0.6  0.3)) 0.6 = 2.03
kN/m1
q fin = 2, 04 kN/m1

L = 5  q fin  L4
250 384 E I

1
E = 11, 000 I= b
N/mm2 h3
12

b = 32 - 59 mm

b = 32 wlimit = u fin b = wlimit = u fin


mm 59
mm

Introduction to Timber Structures 111


L 5 q  L4 L 5 q  L4
fin  fin
250 384 E=  I  250 384 E  I=
384  E  I = L4 384  E  I = L4
250  5 qfin L 250  5 qfin L
384 11, 000  1  32  h34 384 11, 000  1  59  h34
12 =L 12 =L
250  5  2.03 L 250  5  2.03 L
4 4
3 L4439 h= 81843 hL=
L L
3 3
(16, 4  h) = L3 ( 20, 2  h ) = L3
L L
= 16, 4 = 20, 2
h h

Introduction to Timber Structures 112


L L
Rule of h= á
thumb:
15 20

Example

L 3000
Estimate the beam h= = = mm → 46 x 171 mm
height: 162
18,5 18,5

Check the final deflection in Serviceability Limit State (SLS):


1 1
I =  b  h3 =  46 1713 = 1917 104 · mm4
12 12
5q
u =  L4
 fi
n
fin 38 E  I
4
5 2.03 30004
u fin =  = 10.2 mm
384 11, 000 1917 104
L 3000
w = 0, 004  = = 12.0
L=
li
m 250 250
it
u = 10.2 =
Unity fin 0,85 ≤ 1, 00 OK
Check:
wli 12.0
mit

Check of the stresses in Ultimate Limit State


(ULS): kN/m1 (Consequence Class 1)
qd = Gk  g + Qk  q = (0.461.08 +
2.251.35) 0.60 = 2.12
1 1
M =  q  L2 =  2.12 
3.02 = 2.39
Introduction to Timber Structures 113
kNm
d 8 d 8
1 1
W =  b  h2 =  46 1712 = 224 103 · mm3
6 6
M d 2.39 106
 = = = 10.7 N/mm2
d W 224 103

24
f
fm,d m,k =  0.8 = 14.8 N/mm2
1.3
= k m
m
o
d

Introduction to Timber Structures 114


 = 10.7 = 0,
Unity d 72 ≤ 1, 00 OK
Check:
fm 14.8
,d

Wherein is:
I = moment of inertia, in mm4
E = modulus of elasticity, in
N/mm2 b = width of the beam,
in mm
h = height of the beam, in mm
L = span in mm
M = design value of the bending moment, in kNm
d = permanent load, in kN/m2
G
= variabel load, in kN/m2 (floor load = 1.75 kN/m2 variabel lightframe walls
k = 0.50 kN/m2)
Q = representative load for determining the deflection including creep
k = design load, including the safety factors
q = final deflection (including creep)
fin
= maximum deflection
qd
= design value of the bending stresses, in N/mm2
u
= design value of the strength, in N/mm2
fin
wli
mit
d
fm,
d

Introduction to Timber Structures 115


Annex 4: Rules of thumb for Timber Constructions

COLUMNS AND WALLS


Element Horizontal en vertical section Usual height L Critical factors for remarks
between the
d dimensioning
supports

Glulam wooden ≤ 4.0 m1 15 to 20 - compression ( Columns with extensive


column L loads (several floors)
 15 )
d L
needs a lower ratio
- compression and d
L
buckling (  15 )
d

Posts in light ≤ 4.0 m1 20 - 35 - compression and With nailed plywood for


frame walls buckling stability
- thickness for minimal
isolation ( Rc = 4,5

center to center usually 400 to 600 mm m2·K / W)


(standard plywood dimensions)

Sawn wooden ≤ 4.0 m1 15 - 30 Columns with extensive


columns loads (several floors)
L
needs a lower ratio
d

Introduction to Timber Structures 116


FLOORS AND FLOOR
BEAMS

Element section and plan Usual Usual L or Critical factors for


L
thickne span L d h dimensioning /
remarks
ss [mm] [m]
/
height
[mm]

Plywood 12 - 0,3 - 30 - - deflection


floors 30 0,9 40 - concentrated loads

- bending strength

Sawn 16 - 0,6 - 25 - - deflection


wood 25 0,8 35 - bending
floors
strength
With
tongh
and….

Sawn 140 - 2-5 15 - - deflection


wood 286 20 Maximum length: 5,5
floor m1
Center to center: 305 -
beams
407 - 488 – 610 mm
width: b  h
3
Strength
class C18 /
C24 Spruce
Glulam 300 - 6 - 14 -
floor 1000 15 18 and pine
beams
- deflection
width: b  h for
instability
8
center to center
distance approx. 4 to
5 m1 Spruce, larch
and iroko
CLT floors 60 - 4 - 28 -
400 12 32

Introduction to Timber Structures 117


ROOFS AND PURLINS 1
Element section and plan Usual height [mm] Usual span L Critical factors for dimensioning / remarks
L [m] d

Purlins flat roofs 140 - 286 2 - 5,5 18 - 22 - deflection


Sawn wood Maximum length: 5,5 m1
Center to center: 305 - 407 - 488 – 610 mm
h
width: b 
3
Strength class C18 / C24
Spruce and pine

Purlins flat roofs 300 - 1400 6 - 30 16- 18 - deflection


Gluelam h
width: b  for instability
8
L L
center to center distance approx. S  −
3 5
Spruce, larch and iroko
L
pre-arched:  
150

Purlins sloped roofs 140 - 286 2-5 18 - 22 - deflection


Maximum length: 5,5 m1
Center to center: 305 - 407 - 488 – 610 mm
h
width: b 
3
Strength class C18 / C24
Spruce and pine

Stressed skin panels 100 - 250 3-7 30 - 35 - deflection


- supported by purlins

a  300 - 500 mm

Introduction to Timber Structures 118


ROOFS AND PURLINS 2

Element section and plan Usual Usual L Critical factors for


height span d dimensioning / remarks
[mm]
L [m]

Trusses 1200 - 6 - 4 - - Strength of the


without 2000 10 6 connections
- Bending in
girders
the edge
beams Center
to center 
600 mm
- Strength of
Trusses 1000 - 6 - 5 -
with 3000 20 7 the
girders connections
Center to
center  2-
5 m1
- Strength of
Truss 1500 - 12 - 8 -
parallel 3000 25 10 the
connections
Center to
center  4-
6 m1

Introduction to Timber Structures 119


PORTALS

Element front view c.t.c [m1] h Usual Critical factors for


span L dimensioning / remarks
[m]

Glulam beam
on columns

Portal frame

Three
hinged
portal frame
Angled
corners
Three
hinged
portal frame
Arched
corners
Arched frame

Introduction to Timber Structures 120


Annex 5: Standard timber sizes

Sawn Timber

Scandinavian Other
Lumber
Standard
(SLS)
38 x 89 mm 32 x 100 40 x 146 mm 59 x 146 mm 71 x 146 mm
38 x 120 mm 32 x 125 40 x 171 mm 59 x 156 mm 71 x 171 mm
38 x 140 mm 32 x 150 46 x 146 mm 59 x 171 mm 69 x 194 mm
38 x 184 mm 32 x 200 46 x 171 mm 57 x 194 mm 69 x 219 mm
38 x 194 mm 44 x 194 mm 69 x 244 mm
38 x 235 mm 44 x 219 mm 69 x 269 mm
38 x 286 mm

Glued Laminated Timber (Mayr-Melnhof)

Introduction to Timber Structures 121


Cross Laminated Timber

Introduction to Timber Structures 122

You might also like