Вы находитесь на странице: 1из 104

Timber structures - 1

1.1. INTRODUCTION
The twentieth century has already seen incredible technological advances,
some of which have resulted in profound changes in the construction industry.
New materials have been introduced, new manufacturing methods have enabled
traditional materials to be used more efficiently or more cheaply, and the demand
for other traditional materials has declined.
Since ancient times, wood and stone have been important construction
materials. Stone has diminished in importance, but wood is still our most versatile
building material. The use of wood for structural purposes has been criticised as
being expensive, inefficient or perhaps simply old-fashioned, and it would
eventually be eliminated from all construction uses to be replaced by concrete,
metal and plastics. In fact, this dismal forecast of only a short time ago can now be
seen to be fundamentally incorrect. The decline has been halted and it is
progressively replacing its competitors in wide divergent fields. By weight,
construction uses more wood than cement and steel every year.
Anyone, who has used wood, whether in his own home workshop or in
some enormous industrial production process, will be aware of the simplicity with
which it can be worked by hand or by machine. It is, in fact, incredibly simple to
fabricate structures from wood and, even in the most sophisticated production
processes, the tooling costs are relatively low compared with competitive
materials. Wood is ideal if it is necessary to erect an individual structure for a
particular purpose but it is equally suitable for small batch or mass production,
especially now that finger-jointing processes have been largely eliminated. When
these working properties are combined with the other advantages of wood, such as
its high strength to weight ratio, its excellent thermal insulation and the unique
aesthetic properties of finished wood, it sometimes becomes difficult to
understand why alternative materials have ever been considered. Perhaps the
11

Wood as a structural material

greatest contrast between wood and competitive materials arises in connection


with maintenance and damage repairs. A major criticism of a wooden product is
the high cost of maintenance.
There is a tendency for engineers and even architects to reject this ancient
and natural structural material in favour of the more uniform and more modern
manufactured materials but there is no need for them to be suspicious. If they are
given as much information about wood as about these newer materials, such as its
advantages and the correct way in which it should be utilised, perhaps they will
understand it better. They will find that it continues to be a reliable structural
material when many of the others have ceased to be available and have been
forgotten.
However, there is one feature of wood that is unique amongst all structural
materials: it is a crop that can be obtained whereas its competitors such as stone,
brick, metal and plastics are all derived from exhaustible resources. This feature is
alone sufficient to ensure that wood will continue to be used as a structural
material virtually forever and it is also likely to ensure that wood remains the
cheapest of all structural materials, [1].
The variability between woods of different species may appear to be a
disadvantage to the unintelligent user but it is, in fact, a distinct advantage, as
different species have different properties and there is almost always a suitable
wood for a particular purpose. The variability that occurs within wood from the
same species can be tolerated by increasing dimensions to introduce safety factors,
an acceptable procedure in view of the low cost of wood in comparison with
competitive materials. Wood is strong with outstanding rigidity in bending and
strength in compression. It is resilient, attractive and warm to the touch, easily
worked with either hand tools or semi-automatic industrial machinery. It is easily
joined with simple fixings such as adhesives, nails or bolts which do not require
elaborate tools. Wood has exceptional stability in the longitudinal direction, even
when subjected to fluctuating moisture content or exceptional temperatures. Wood
is normally durable when dry and comparatively inexpensive processes enable it
to withstand the most destructive biological agencies, such as insects and fungi.
Moreover, wood is free from corrosion. Its cellular structure ensures good
insulation properties and, while it is combustible, it maintains its integrity when
exposed to serious fires in which metals melt and concrete disintegrates.

1.2. WOOD STRUCTURE


1.2.1. Anatomy of wood
Features of wood structure include those of the wood-forming substance
and those of the whole wood. As a plant a tree consists of a crown of branches
with leaves, generally supported on a single main stem known as the trunk (or
12

Timber structures - 1

bole) which connects the crown to the roots in the ground (fig.1.1.a). The trunk
grows outwards around a leading shoot by adding new rings of timber. Usually
one ring is added each year, but as this is not always the case it is better to refer to
the rings as growth rings rather than annual rings. The more rapid the growth, the
wider the growth rings, and when trees of a particular species are compared, the
wider the growth rings, the less dense and strong is the timber.
A further annual growth is accompanied at the
inner surface of the sapwood by the death of
cells and their conversion into heartwood.
The only significant difference between
sapwood and heartwood is the large amount of
material that is deposited in the latter,
apparently waste material arising from the
living processes of the tree. These deposits in
the heartwood cells are often significantly toxic
so that the heartwood is generally more
resistant to insect and fungal attack than
sapwood. These deposits also tend to make
heartwood more stables so that it is much more
(a)
resistant to swelling and shrinkage with
changes in moisture content.
Heartwood (fig.1.1.b) is the older wood in
the central portion of a tree, which has ceased
participating actively in the physiology of tree
life.
Sapwood is the newer wood, which
usually appears as a lighter coloured band
immediately within the bark, extending inward
from a few too many annual rings, depending
heartwood
upon species. The structural properties of wood
from either sapwood or heartwood regions
sapwood
(b)
used as structural lumber or plywood veneer
are so similar that no distinction is required.
Figure 1.1
- Tree components and Heart and sap portions do have different
permeabilities, however, and sapwood is more
transverse cross-section
amenable to penetration by wood preservatives
and fire-retardant solutions.
The basic properties of a piece of wood depend upon the species of the tree
from which it is derived. Woods are commonly divided into softwoods (the conebearing plants that are conifers) and hardwoods (the broad-leaved plants meaning
dicotyledonae and monocotyledonae). The terms hardwood and softwood are only
13

Wood as a structural material

vaguely related to the actual hardness of the wood, since some softwoods are
actually harder then some hardwoods. The cheaper hardwoods approximate in
cost to the more costly softwoods. Typical softwood species are the pines, firs,
spruces, and redwoods, while typical hardwood species include the oaks,
maples, beeches and birches.

1.2.2 Methods of conversion


The first level of wood for construction is the log. The logs are converted
into sawn wood by means of conversion saws. The cutting of logs into sections
before seasoning is known as conversion. Subsequent re-sawing and shaping is
commonly called manufacture. Basic methods of cutting are sawing; peeling for
producing plies for plywood; slicing of thin decorative veneers and cleaving or
splitting, as used generally before the development of tempered steel saws and
now for palings.
Usually the first cuts are on either side of a log to give a centre block and
sideboards. Rollers to the edging saws often divert these sideboards, although they
are largely wasted and in some mills they are used for the manufacture of particle
board or paper pulp. The centre block is turned on one of the cut sides and passed
through a second frame saw to produce deals or batten. The manner in which the
log is sawn is usually considered to be relatively unimportant. It is not so, because
the manner can influence the behaviour of the sawn wood. The simplest technique
is to make a large number of parallel cuts, a method known as through-andthrough, flat sawn or back sawn (fig. 1.2.a). The outer boards are largely cut in the
tangential plane while the middle board is in the radial plane, the angle between
the annual rings and the surface of the board progressively varying through the
intermediate boards.
Another manner is quarter-sawn boards (fig. 1.2.b). It is particularly
suitable for use as flooring as they do not suffer the cupping that is a characteristic
of tangential or outer flat sawn boards.
An alternative method of conversion is to make three through-and-through
cuts (fig. 1.2.c) to provide two flat sawn boards from the centre of the log. These
will naturally include any heart defects, which can then be removed when the
boards are re-sawn. The remaining wood consists essentially of two half logs
often known as wainscot billets. These billets are then turned on to their flat face
and re-sawn to give a number of boards.
Figure 1.3 shows three ways in which the direction of growth rings can
relate to surface: i.e., as end grain; more or less parallel to surfaces in plain sawn
boards and at right angles to the surfaces of radial sawn boards. Radially sawn
boards shrink less in their width and are less liable to cup and twist; they are easier
to season and wear more evenly than plain sawn boards. Unfortunately however,
14

Timber structures - 1

methods of cutting that produce a high proportion of quarter-sawn timber are


wasteful and the extra cost is justified only where advantages are important.

(b) quarter cut

(a) through and


through sawn

(c) billet sawn

Figure 1.2 Methods of log conversions

end grain
(transverse section)

flat sawn
plain sawn
slash sawn
(tangential surface)
rift sawn
quarter sawn
(radial surface)

Figure 1.3 Surfaces of cut timber


The transformation process of a tree in wood products basically includes
the following stages: trunk, log and sawn wood.
15

Wood as a structural material

is barked (peeled)
TRUNK

is converting
LOG

SAWN WOOD

Sawn wood products are divided into centre block and sideboards.
Sideboards are the outer parts of the log and they are used for non-structural
wood products. Centre block is also sawn in different manner for structural
products.
The structural products obtained through successive steps of wood
conversion are:

log

- planed square edged board


- planed tongued and grooved board
- planed tongued and grooved with V joint board
(match boarding)
- boards
- plain weather-board
- rebated weather-board
- ship-lap weather-board
- sawn timber columns
- sawn timber beams

Non-structural wood products include:


- floorboards

- joinery (millwork)

- doors
- door frames
- door stops
- architrave
- skirting
- panelling
- windows
- window frames
- surrounds and faces
- large-boards and cladding

The wood elements with square section are well-defined by the following
terms (fig. 1.4):
Face

- faces and edges;


which have:
- width and thickness.
16

Thickness

Edge
Width

Figure 1.4 Wood element terms

Timber structures - 1

When wood products are used for building they are called sawn lumber.
Structural calculation is based on the standard net size of a piece of lumber.
Most structural lumbers (wood sawn) are dressed lumbers, and this means that
the lumber is surfaced to the standard net size. The standard net size is smaller
than the nominal size. Lumber is dressed on a playing machine for the purpose
of obtaining smooth surfaces and uniform sizes. Typically lumber will be S4S
(surfaced four sides) but other finishes can be obtained (e.g. S2S1E - indicates
surfaced two sides and one edge). The designer may have to allow for shrinkage
when detailing connections, but standard dimensions are accepted for stress
calculations (fig. 1.5).

Nominal size
Standard dressed (net) size

Figure 1.5 Wood sizes

Usually, a sort of sawn wood called timber is used in buildings. Timbers


are larger-size wood members, known as timbers, because of their large crosssectional dimensions are not produced in a dry condition. An excessive amount
of time is required to season these members to a dry condition. The resistances
have been adjusted to account for the higher moisture content of timbers.
Large timbers are commonly rough sawn to dimensions that are close to
the standard net sizes. The textured surface of rough sawn lumber may be
preferred for architectural purposes and may be specially ordered in smaller size.
The cross-sectional dimensions of rough-sawn lumber are larger than the standard
dressed size. A less common method of obtaining a rough surface would be to
specify full-sawn lumber. In this case, the actual size of the lumber should be the
same as the nominal size.
Historically, the size of trees in the forest determines the size of the timber
that may be produced. One hundred years ago timber with cross-sections of
150x450 mm and lengths up to 20 m were commonly available. Today, in many
countries, timber over 75x225 mm and more than 5 m long attracts a cost
premium due to scarcity. However, if larger sizes are required, several timber
members can be combined to form a composite member. Therefore, related to the
tree size, there is a number of standard timber products (or lumbers). The main
categories of lumbers are boards, dimensions, and timbers. They are also divided
into a large number of subcategories presented in figure 1.6.
17

Wood as a structural material

SAWN LUMBER
(SIZE CATEGORIES)

BOARDS

LUMBER
(DIMENSIONS)

Light framing

Joist
Plank

Decking

TIMBERS

Beam
Stringer

Post
Timber

Figure 1.6 Wood products classification


The last five divisions shown above are the most utilised structural lumbers.
Their dimensions and specific nomenclatures in the Romanian code are:

h = thickness
(height, depth)

1. timber slat (strip)


h 40 mm
b 60 mm

b = width

2. boards (lumber)
h 40 mm
b 80 mm
h

h = thickness
(height, depth)

b = width

3. sawn timber collar beams


40 mm h 100 mm
b 80 mm
4. thick plank
40 mm h 100 mm
b 100 mm

Figure 1.7 Sizes of wood products


5. sawn timber columns and beams
h 100 mm
b 100 mm

1.3. NATURAL DEFECTS


Wood has various natural defects, which can influence the strength and
thereby arrive at a value, which is acceptable for these defects. Defects may be
classified as natural defects, chemical defects, conversion defects and
18

Timber structures - 1

seasoning defects. All the defects may degrade wood, with the degree of
degradation being reflected in varying degrees of loss in mechanical properties.
The natural defects, which influence the structural behaviour of wood
elements, are:
- Knots
A knot is the part of a branch, which became enclosed in a growing tree.
Where the fibres of a branch are completely continuous with those of the tree a
live knot results, and where the fibres are continuous with those of the tree to
the extent of at least three-quarters of its cross sectional perimeter, the knot is
inter-grown. A dead knot has fibres inter-grown with the surrounding wood.
Dead knots are sometimes tight but they are often loose.
Knots attacked by fungus are termed unsound or decayed.
The weakening effect of a knot is brought about by
the local disturbance in the grain direction it produces and it
is not due to any inferiority in the material of the knot.
knot (local disturbance of grain)

- Grain defects
These are the measure of the deviation of the fibres from
the longitudinal axis of the piece. If fibres occur at an
angle then any forces applied along the longitudinal axis
will create components of force on those fibres, thus
reducing strength. Grain defects can also occur in the form
of twisted-grain, cross-grain, flat-grain and spiral-grain, all
of which can induce subsequent problems of distortion in
use.
- Annual ring width
This can be critical in respect of strength in that excess
width of such rings can reduce the density of the timber.
Although this type of defect is not as important as other
features but there does the average number of growth rings
per 25 mm indicate some limits.
growth rings

- Fissures and cracks


A fissure is any separation of fibres in a longitudinal plane
and includes checks, shakes and splits. Their existence
reduces the cross-sectional area, resisting shear and
bending stress.
19

Wood as a structural material

- Fungal decay
Wormholes are permitted to a slight extent provided that
there is no active infestation. Wood wasp holes are not
permitted. Decayed wood should not be accepted.
Next to knots the commonest causes of degrade in timber are seasoning
defects. These defects are bowing, springing, twisting and cupping. Seasoning
defects are directly related to the movement that occurs in timber due to changes
in moisture content. All such defects have an effect on structural strength as well
as on fixing, stability, durability and finished appearance.
Chemical defects may occur in particular instances when timber is used in
unsuitable positions or in association with other materials. Most woods are
slightly acidic and produce acetic acid if stored in damp conditions. Timber such
as oak contains tannin, which corrode metals. Gums and resins adversely affect
working properties and ability to take glue and surface finishes, while silica in
some hardwoods blunts tools. Conversion defects are basically due to unsound
practice in the use of milling techniques or to undue economy in attempting to use
every possible piece of timber converted from the trunk. The most important
defect of this group is the wane:
- Wane
This is a reduction in the cross-sectional area of the
rectangular timber section across the corners due to the
section being taken from a location close to the outer
circumference of the tree.
one or more corners affected

1.4. WOOD PRESERVATION


It must be accepted that wood is perishable. Unfortunately, the various
wood destroying insects and fungi are unable to distinguish between forest waste
and wood in useful service. In the past, wood was used mainly in forest areas
where it was readily available. Degradation was accepted but replacement was
comparatively simple and inexpensive. In contrast, wood is now a valuable
commodity, transported considerable distances between the production forest and
the ultimate user. While fire and moisture content changes must be clearly
recognised as causes of degradation in wood, it is, of course, the various living
organisms, both microbiological and animals, which are perhaps the most
important.
It is possible to find woods with natural resistance to almost all-destructive
organisms, the heartwood generally being more durable than the sapwood.
20

Timber structures - 1

However, physical properties, availability and cost may preclude selection in this
way. Success in preservation depends on the timber species, the size and condition
of the specimen, the effectiveness of the preservative and the resulting depth of
penetration and the amount, which is retained.
In buildings in particular, structural precautions to prevent wood from
becoming wet will be sufficient to avoid the major hazards of fungal attack and
associated insect infestation. In fact, the most important and accessible
preservative action is the method of painting. It works as a barrier, isolating the
wood from attacking insects or fungi. But minor imperfections or damage to the
paint film will permit absorption of water while the remaining paint protection
will simply restrict evaporation and thus allow dampness to accumulate.
Joinery frequently has only a single primer coat as protection on the hidden
faces, despite the fact that these are precisely the areas where most damage occurs
during installation and where contact with adjacent damp materials may result in
moisture absorption.
Minor damage may permit access to the wood and protection is clearly
more efficient if the wood is impregnated in depth rather than simply coated. This
is the origin of pressure treatment with coal tar, but all treatments of this type
cause a fundamental change in the appearance of the wood. Such changes may be
aesthetically unacceptable and unsuitable for use in many situations, if the
treatment is considered dirty. The alternative is a complete abandonment of the
barrier principle in favour of a toxic action, low retention of highly toxic
compounds achieving preservation without pronounced alterations in the physical
properties and appearance of the treated wood. This principle is applied in the
form of the salt preservatives. A variety of multi-salt preservatives have now been
developed. These multi-salt preservatives are giving excellent service throughout
the world but they suffer from two distinct disadvantages. They must be applied to
wood at low moisture content in order to achieve the necessary absorption of
preservative solution, yet they result in treated wood of exceptionally high
moisture content which must be reduced before it can be used for many purposes.
The changes in moisture content can result in severe distortion and possibly a high
rejection rate. The cost of solvent alone is sufficient to ensure that organic solvent
preparations are more expensive, yet they are progressively replacing the aqueous
multi-salt preservatives wherever these factors are of importance, as in the
treatment of joinery (millwork). Vacuum and pressure impregnation treatments
may provide improved penetration.
In some cases the cost advantages of preservation are obvious, for example
non-durable timber telegraph posts which have been impregnated with
preservative. Good construction in buildings should not normally put even
perishable timbers at risk, but with the smaller sizes and increased sapwood
content of structural timber today, a cost of preserving them of less than one per
cent of the total cost of building a typical house is justified. Table 1.1 presents the
categories of needs for preservation of structural timbers:
21

Wood as a structural material


Table 1.1
Condition of use
1. Negligible risk, even for timbers with low
inherent resistance to biological degradation
2. Low risk or where remedial action is simple
3. Where experience has shown that there is an
unacceptable risk of decay
4. Where timber cannot be protected by design
from a continually hazardous environment

Preservation
Unnecessary
An insurance against the cost of repairs
Necessary
Essential

The following preservatives are recognised in the standards:


- Preservative oils:
Creosote
Creosote-coal-tar solutions
Creosote-petroleum solutions
- Oil-borne preservatives:
Pentachlorophenol
Copper naphthenate
- Water-borne preservatives:
Chromated zinc chloride (CZC)
Fluor chrome arsenate phenol (FCAP)
Tanalith (Wolman salts)
Celcure
Chemonite
Greensalt (Eradlith)
Boliden salts
The selection of the proper method of treatment is as important as the
selection of the proper preservatives. The best preservatives known will not
increase the life of wood appreciably if it is not effectively injected into or painted
on the wood.

1.5. FIRE RETARDANTS


Fire will only occur when combustible material is subjected to sufficient
heat in the presence of oxygen. In the absence of any of these three components
ignition cannot occur.
Heavy timber construction has traditionally been recognised to provide a
fire-resistance building. This is primarily due to the large size of the members, the
connection details, and the lack of concealed spaces. Such a construction type has
often satisfied the fire-resistive requirement in all codes by simple prescription.
22

Timber structures - 1

The fire endurance rating R, in minutes, is determined for beams or


columns exposed to fire on either three of four sides. It is expressed by the
following equation:
[min.]
(1.1)
R 2.54 ZbG
where: - Z = factor dependent on load applied and member type. It has values
between 1 and 1.5;
- b = width dimension of cross section of beam or of larger dimension
of a column before exposure to fire;
- G = beam or column cross-sectional factor.
Within a building it is necessary to have fire resistance. This is the ability of
the building components to withstand fire penetration so that an accidental fire
remains isolated for a sufficient period to permit the occupants to escape and to
give reasonable time for the fire service to arrive and prevent further damage.
There are two principal systems, which are used to increase the fire
resistance of wood structures and to reduce the surface spread of flame. The
simplest system is perhaps the application of a special intumescent paintinsulating barrier, which prevents rapid increase in the temperature of the wood
beneath and contains fire-retarding gases, which further restrict the possibility of
wood ignition. Intumescent coatings are only realistic where they can be applied
as part of a decorative system.
The alternative is the impregnation of the wood with fire retardant salts.
These prevent normal ignition, although some surface charring may occur to a
very shallow depth. These impregnation treatments are particularly suitable for
structural sawn wood. Salts such as sodium tetraborate, diammonium phosphate,
trisodium phosphate, diammonium sulfate, and salts of boric acid have long been
used as fire retardants. Many fire-retardants salt treatments are very hygroscopic.
Accordingly, most formulations are not recommended for use where relative
humidity is over 80%. Recently fire-retardant resin treatments have been
developed.

1.6. WOOD AXES


Wood is a biological material and it is immensely varied in its structure.
Because of its internal structure, wood is orthotropic, which means that it is
having different properties in each of the three principal directions. The three
structural axes of wood are longitudinal, radial and tangential and are
designated in accordance with their orientation in the growing tree (fig. 1.8).
The top face is a transverse cross-section showing the curved concentric
annual growth rings. This face may be designated the RT = Radial -Tangential
23

Wood as a structural material

plane as it is defined by the radial and tangential axes of growth symmetry. These
side-grains may be tangential either as shown in the left face (LR = Radial Longitudinal plane) or in the back face (LT = Longitudinal Tangential plane).
Usually the sides of pieces of wood are surfaces intermediate between the LT
plane and RT reference planes. The main axes of the wood are required to define
mechanical properties or other characteristics, which are used in wood design.
Reference to this nomenclature will be made wherever the mechanical properties
are influenced by the orthotropy of the wood. Strengths and modulus of elasticity
vary in the three directions, and there are six values of Poissons ratio. Shrinkage
(or swelling) occurring as woods moisture content changes also differs in the
three directions; this is what may cause wood to warp as it either dries out or takes
on additional moisture.
Longitudinal axis

Tangential axis

Radial axis

Figure 1.8 Orthotropic axes in wood


The non-isotropic nature is probably the most serious characteristic with
which the designer of wood structure must cope.

1.7. LUMBER GRADING


Grading is the process of classifying lumber according to quality for a
particular use. Grading is accomplished through rules that define both desirable
characteristics and limits to undesirable characteristics of the lumber, and provide
methods of quantifying the undesirable characteristics.
The purpose of grading rules is to maintain a standard among lumber mills
manufacturing from the same or similar species, so that the quality of the lumber
for some particular use can be defined and controlled, regardless of the overall
character of the logs from which the lumber is produced. Thus, lumber of a
particular grade from different mills will have essentially the same strength and
stiffness properties. The structural designer can specify lumber of a particular
species group and grade, and can be reasonably assured that the lumber received
24

Timber structures - 1

will be suitable for the intended structural use, regardless of which mill produced
it. The grade stamp placed on each piece of lumber will define the quality of
material to the user. Grading rules may differ according to the size of the piece
being graded. So grading is related to both size and use.
Structural designers are interested in strength and stiffness, so modern
grading rules provide for what is sometimes called stress grading.
The two methods used for stress grading are:
- visual grading;
- machine grading.
In visual grading of lumber for structural uses, the grader examines each
piece to determine the type, location, and size of various defects that might affect
its structural strength. Then, according to rules that quantify the effect of each
defect, the piece is assigned to a grade. The grade name or symbol is stamped on
the piece, giving the user a means of determining the probable strength and elastic
modulus of the material.
Machine grading is used to determine lumber properties by non-destructive
testing. It is based on the principle that all strength properties bear some
relationship to the modulus of elasticity, E. In machine grading, each piece is first
subjected to visual grading. Then it passes through the grading machine, which
bends the piece to a predetermined curvature and measures the forces required for
its bending. Deflection and force required are then used to determine the modulus
of elasticity. Allowable stress levels are determined from the E value.
The minimum requirements for visual grading standards have been laid
down in the European Code EN-518 Structural timber Grading Requirements
for visual grading standards. Requirements for machine grading can be found in
EN-519 Structural timber Grading Requirements for machine strength graded
timber and grading machines.
Guidance on the use of timber in building and civil engineering structures is
given also by the Romanian code STAS 857-83. Structural timber elements are
divided in quality classes according to the load type, number and size of defects.
Table 1.2 presents the main quality categories of timber elements according to the
presented code rules.
Table 1.2
Quality class
I
II

III

Load type and destination


Timber elements are subjected to tension and bending
(Truss girders, beams and wood dowels)
a) Timber elements are subjected to compression and bending
b) Timber elements are subjected to tension and tension + bending
where effective stresses are 70% of allowable wood strengths
Secondary timber elements (Roof covering)

25

Wood as a structural material

The requirements to be attained by each quality class of timber elements are


presented in tables 4 & 5 of STAS 857-83. In figure 1.9 are presented the
requirements related to the knots presence.

d1

d2

d3
l

20 cm
d1

50 cm

I:

d1

d2

d3

b
4

20 cm

II :

d1

d2

d3

b
3

40 cm

d2
d3

d1

d2

20 cm

d3

III :

d1

d2

d3

b
2

Figure 1.9 Quality categories of timber elements

1.8. TIMBER CONSTRUCTIONS DEVELOPMENT


1.8.1. Timber frames for houses
There are two important challenges for those who build timber-frame
homes today. Both should concern all those associated with the building project,
for timber-frame house design and construction processes. They are more like a
chorus line than a solo dance, requiring all parties to play an active role.
The first challenge is to cherish and nurture the values and standards set
centuries ago, to understand in both spirit and substance the legacy of durable,
classic timber-frame building that we have inherited from our forefathers. What
is it that makes a structure so fine that it survives numerous centuries? Is it the
construction, the architecture, or both? What makes people love a building
enough to repair it and restore it? To give direction to our modern efforts, it is
necessary to investigate the historical precedents first. To know where we shall
go, we need to know where the craft has been, [3].
The second challenge is to bring the timber frames into the 21st century,
which requires a careful analysis of the limitations and opportunities inherent in
the building system. In this chapter well look at both challenges, beginning
with an exploration of the development of the timber constructions through
history and the attributes that give it timelessness.
26

Timber structures - 1

Figures 1.10.a, b & c illustrate the


starting point and the beginning of wood
constructions, from which all later forms up to
the present day have arisen.
Figure 1.10 Primitive structures
(a) - Primitive roof

(b) Round tent

( c) Long tent with ridge purlin

The first timber frame


buildings appeared at about
the time of the birth of Christ.
They probably evolved from
skeleton of tied-together poles
around which skins or
mattings were wrapped. Such
dwellings were usually the
homes of nomadic peoples
(fig. 1.10.b, c).
(a) Wood structure

Figure 1.11 Early wood structure


When permanent settlement,
became desirable, pole
(b) Sharpened
skeletons slowly gave way
wood post in ground
to more stable and durable
timber structures as presented
in fig. 1.11.a.
The early wood structure was made of pieces lashed together and
earthbound posts. The posts (fig.1.11.b) of the house were stuck directly in
holes and stabilised by compacted earth, a technique that allowed a minimum of
interconnection between timbers and caused the post to rot. Rigidity arises from
ground anchoring of the frame.
27

Wood as a structural material

(a) timber house frame

Figure 1.12 Timber house frame with early mortise-and-tenon joint

(b) early mortise-and-tenon joint

The creation of the mortise-and-tenon joint


(fig.1.12.b) between 500 BC and 200 BC, which meant
the tools and technological sophistication, existed to
work the wood, allowed a new type of building frame,
fig.1.12.a. The joint, bringing frame out of ground, gave structure a longer life.
It also required early builders to make stronger joints and develop structural
resistance to loads.

Figure 1.13
- Log home

The development of timber-frame buildings differed from region to


region, depending on many factors, such as the availability of materials. In
Egypt, Persia and Greece, where timber was limited, early temples were built
28

Timber structures - 1

with wooden columns and lintels; their forms were the basis for later stone
temples. By contrast, in parts of northern Europe that were heavily wooded, a
building style emerged around 9th century in which the timber frame was
vertically infilled with thick wooden planks or more timbers, making a veritable
fortress of wood. Also in heavily wooded areas, timber framing was
occasionally used with log building. For example, fig. 1.13, some Swiss chalets
have horizontal logs laid between the joined timbers. Stacking logs is a more
basic means of creating a structure, yielding a massive and earthbound look.
Where log construction was dominant, timber framing was still used for the roof
system, fig. 1.14. Timbered buildings seemed to rise out of ground with more of
a vertical statement and to demand use of complex joinery and geometry, [3].

Figure 1.14 Timber frame home

Figure 1.15 Timber frame structure of the Middle Ages


29

Wood as a structural material

By the Middle Ages, the craft of timber framing was fully developed, and
by the standards of the time, timber-frame house required the least amount of
labour. Timber was therefore preferred, even when it was relatively scarce, and
the condition and variety of available trees often determined the shape and the
style of the frame, fig. 1.15, [3]. Some of the greatest carpentry of all times is
evident in timber-framed roofs of Middle Ages. During this period, English
carpenters developed ingenious systems for spanning great widths without long
timbers. As European forest resources became depleted, masonry was
commonly used for exterior walls.
A typical English timber-frame house of this period is presented in
fig.1.16.
Early American timber-frame houses also shared a similar design, derived
from generally accepted notions about function that had evolved through many
centuries, and an example is presented in fig. 1.17.
Today building timber-frame structures to the highest standards requires a
broad understanding of theory and good practice combined with some creativity
and flexibility. Each frame starts out as not much more than a blank picture
made up of the hopes and dreams of the future occupants brought down to earth
by budget limitations and site requirements, [3].

Large, naturally
curved braces

Figure 1.16 - Old English style timber frame


30

Timber structures - 1

Figure 1.17 Typical American timberframe house


Timber framing has more in common with designing and making wooden
furniture than it does with conventional building procedures. In the same way
that a handcrafted chair is more than a place to sit, a timber frame is more than a
structure. In both cases it is virtually impossible to separate design from
construction. Be it chair-making or timber framing to make successful decisions
about woods, finish, style, function, and construction, you must first understand
the structural requirements of the craft. To get the necessary overview, examine
typical engineering considerations and the physical characteristics of the timber
frame.

1.8.2. Timber frames for bridges


Bridges built either wholly or partly of timber include pedestrian,
highway, and railway bridges and special types such as pipe bridges. They
include trestles, simple girder, arch or truss bridges, and even suspension
bridges. Although wood is an excellent material for many types of bridge, its
value for bridges is often not clearly recognised, perhaps because of a fear of
either impermanence or weakness. However, properly constructed and
maintained wood bridges can be durable. The useful life of modern glulam
bridges is estimated to be as high as 50 years, whereas for steel or concrete
bridges (because of corrosion) it may be much lower. Timber bridges are not
31

Wood as a structural material

necessarily weak; they can carry very large loads, such as heavily loaded
logging trucks.
Timber is a popular material for pedestrian bridges. In these, the entire
superstructure main girders and deck are wood. The pedestrian bridges
starting point is illustrated in fig. 1.18. All-later forms (fig. 1.19), up to the
present day, have been developed from the ancient solution having the same
purpose.

Figure 1.18 Natural pedestrian bridges

Figure 1.19 Primitive timber bridges

32

Timber structures - 1

The development of timber-frame bridges differed from region to region,


depending on many factors. In Europe, which was heavily wooded, a bridge
composite style emerged BC in which the timber frame was the superstructure
and stone or masonry were used for the infrastructure, fig. 1.20. The materials
gave a longer life to structure because they allowed stronger joints and develop
structural resistance to loads.

Figure 1.20 - The Drobeta-Turnu Severin bridge designed by Apolodor and


built by Romans
The selection of wood over other materials is often made on the basis of
aesthetics, but economy, durability, and ease of maintenance also play important
roles in determining the choice. In figures 1.21.a, b, c are shown three examples
of timber bridge structures.

(a)

Figure 1.21 Different


timber bridge structures

(b)

(c)
33

Wood as a structural material

In railway trestles (fig. 1.22.a, b), [18], treated wood piles are frequently
used for the foundation, and treated timbers are used for all other structural
parts, including the cross ties.

(a)

(b)

Figures 1.22 Timber railway trestles


Highway bridges can be built using sawn timbers for the main
longitudinal members. When glued laminated (glulam) girders are used, the
practical span limit is higher. Wood is used frequently for the towers, stiffening
trusses, and floor systems of suspension bridges. Arch bridges have become
practical with the development of glulam. The decision to use wood is often
based on the appearance of the bridge and the blending of its appearance with
the surroundings.

1.8.3. Great timber structures


Over the centuries, man learned to use wood effectively, developing rules
of thumb to aid in planning and building wood structures. These rules were
learned gradually, based only on experience both successful and unsuccessful.
Using these rules, skilled artisans were able to construct in wood, producing
durable, long lasting structures of utility, often of great beauty, and almost
34

Timber structures - 1

always of adequate strength and serviceability. The following pictures present


some examples of the most beautiful timber structures developed over the
centuries. The selected timber structures are great due to both their beautiful
appearance and strength and manner of construction, [1, 3].

Figure 1.23 Church of St. Pierre,


Libresville, Gabon, 1990

Figure 1.25 Dome Timbering Church


of the Sorbone, Paris, France,
Figure 1.24 Spire and transept roof
1655-1742
Cathedral of Notre Dame de Paris,
France, 1165-1250

35

Wood as a structural material

Try to
answer
these

Questions for you

1. Describe the advantages of wood as a structural material.


2. Describe the disadvantages of wood as a structural material.
3. Define the following terms: heartwood, sapwood, hardwood, and softwood.
4. Explain the orthotropy of wood and produce a sketch of wood axes.
5. List the wood products used in building structures.
6. Describe the wood defects and produce sketches of them.
7. Describe what is meant by lumber grading.

36

Timber structures - 2

2.1. INTRODUCTION
The arrangement of fibres in wood, with the long axes of the fibres parallel to the
axis of the trunk, suggests that wood may have different characteristics in the
various directions within itself. Specifically wood is considered to be orthotropic,
having unique and independent properties in the direction of three mutually
perpendicular axes. The mechanical properties of wood used in design process of
a building element are usually referred to the following axes: longitudinal axis
and transverse axis, fig. 2.1. The transverse axis is used instead of tangential or
radial axes because the variability of the same properties about them is less and of
minor importance in timber element design.
R = radial axis

L = longitudinal axis

L = longitudinal axis

T = transverse axis

T = tangential axis

Figure 2.1 Wood design and orthotropic axes


37

Wood properties

Very many properties or qualities of materials have to be considered when


choosing a material to meet a design requirement. In Table 2.1 a property
classification of interest for civil engineers who intend to use wood as a building
materials is given.
Physical properties

Mechanical properties
Manufacturing properties
Economic properties
Aesthetic properties

Table 2.1
Density. Moisture content. Hardness and
Toughness. Electrical properties. Acoustical
properties. Thermal properties. Behaviour in
fire. Resistance to corrosion and environmental
factors.
Strength properties. Elastic properties. Fatigue
strength. Fracture toughness.
Ability to be shaped by machines. Ability to be
joined by adhesives.
Processing cost. Availability.
Appearance. Texture and ability to accept
special finishes.

2.2. PHYSICAL PROPERTIES


2.2.1. Hardness and toughness
Sawn-wood is frequently described in terms of its hardness and toughness,
but these are terms that are difficult to define. Sometimes wood is said to be very
tough because it is difficult to saw or plane, or it has good resistance to abrasion,
splitting or shock loads.
In fact, the ability to resist excessive shock is probably the property that can
best be described as toughness. The meaning of hardness is the ability of wood
to resist penetration, usually measured by the force required for a standard ball to
penetrate for half its diameter. Hardness decreases with increasing moisture
content so that it is necessary to standardise the moisture content, perhaps at 1215%, when carrying out tests to compare various samples, [26].
Occasionally hardness or toughness is used to describe the ability of wood
to resist wear or abrasion, so that hard or tough woods such as hornbill or apple
were used for silent cogs, and maple is used for dance floors.

2.2.2. Thermal properties


Temperature affects both dimensional stability and strength of wood. Wood
expands as its temperature increases, as do other construction materials. Its
coefficients of expansion vary with direction, being largest radially and
38

Timber structures - 2

tangentially, and least longitudinally. Wood is a good insulator, that is, it has a
high resistance to heat flow.
Specific heat is the term used to describe the amount of heat energy that is
required to raise the unit mass of the material through one degree of temperature.
The specific heat of wood is comparatively high, four times as high as that of
copper, but this relates to the mass of material. The density or specific gravity of
wood is very low so that the specific heat per unit volume is also very low
compared with competitive materials such as metals and concrete. However, the
specific heat of wood is of minor importance compared with its thermal
conductivity.
The thermal conductivity of wood is approximately 0.4% of that of steel
and 0.05% of that of copper, having the same order of value as cork and gypsum
plaster, which are commonly used for insulation purposes. Its conductivity k is
0.144 W/mK.
The thermal conductivity varies approximately in proportion to density,
being lowest for low density wood and highest for high-density woods. The
lighter species are obviously the best insulators. Thermal conductivity also
increases significantly with moisture content and it also depends on the type and
location of defects. For practical purposes no distinction is made between thermal
conductivity in the T (tangential) and R (radial) directions. Thermal conductivity
in the longitudinal direction (L) is 2.25 to 2.75 times the value given above for the
T or R directions.
However, generally, the low specific heat per unit volume and the low
conductivity combine to ensure that wood feels warm to the touch and, in
structural applications, wood claddings or floor provide excellent thermal
insulation.
The longitudinal coefficient of thermal expansion, while not related to
density, has been found to differ among various species. The coefficients, which
have been published in the literature on native wood range from 3.06 x 10-6 to
4.5 x 10-6 / oC. A coefficient of 3.6 x 10-6 / oC would be a good value for general
estimating purposes for the more commonly used structural species.
Radial and tangential coefficients of thermal expansion are related to
specific gravity in some technical reports. The values based on specific gravity are
approximated by the following equations:
- for softwoods:
T

= 1.8 x 45 x specific gravity x 10-6

[/ oC]

= 1.8 x 31 x specific gravity x 10-6

[/ oC]
39

Wood properties

- for hardwoods:
T

= 1.8 x 32 x specific gravity x 10-6

[/ oC]

= 1.8 x 32 x specific gravity x 10-6

[/ oC]

Compared to metals the longitudinal thermal expansion of wood is


significantly bigger than the aluminium value (2.3 x 10-6 / oC), and lower than
steel value (1.17 x 10-5 / oC). For this reason, consideration must be given to the
differential thermal expansion of materials used in conjunction with wood.
Because wood is a good insulator, it does not respond rapidly to any
temperature change of the environment; its thermal expansion greatly lags the
temperature changes of the surrounding air, and the ends of members projecting to
the parts that are in the heated (or cooled) interior.
Moisture content changes often have a more important effect on wood's
dimensions than the expansion and contraction effects of temperature.

2.2.3. Electrical properties


The designer of wood structures may not need to carry out electrical design
computations, but he/she will need to know how wood functions as a conduit for
electricity. The resistance of wood to the flow of direct current or low-frequency
current is similar and at moisture contents, which prevail in most covered
structures - quite high. Wood at a low moisture content is normally classified as an
electric insulator, or dielectric, rather than as a conductor
Tangential and radial resistance exceed longitudinal resistance for wood.
Wood density, moisture effects and temperature have effects on resistance of
wood to electrical current.
Metallic salts, such as those used in preservative and fire-retardant
treatments, may lower the electrical resistance of wood considerably. Use of wood
containing such salts should be avoided for applications where electrical
resistance is critical and in processes involving dielectric heating. Electric
moisture meters may give erroneous reading for such wood.

2.2.4. Acoustical properties


The acoustical behaviour of wood will concern the engineer primarily in
relation to the control of noise in structures. These properties are determined by its
sound insulation and sound absorption abilities.
40

Timber structures - 2

Sound insulating values are related to the sound transmission. The


reduction of sound in its transmission through a material is dependent upon the
weight of the material. Since wood has a lower density than many structural
materials, its effectiveness in blocking transmitted sound is not high. Like most
common construction materials, wood alone does not provide good insulation but,
when combined with other materials in typical constructions, it will provide a
structural unit of satisfactory sound-insulating ability.
Sound absorption coefficient for a material is used to determine the total
magnitude of the absorption property of the material. Wood surfaces are fairly
effective as absorbers of incident sound energy. Sound absorbing ability is
expressed as percent absorption of incident energy, and data is available from
numerous sources. Porous and low-density materials are often superior
absorbers.
Wood absorbs low-frequency sound somewhat better than high-frequency
sound, and is equal or superior to many commonly used materials.

2.2.5. Density and specific gravity


Specific gravity (G) or relative density is the weight of a substance to that
of an equal volume of water. Density is the mass per unit volume normally
expressed as kg/m3. Basic specific gravity is defined as:
G

0g
w

where

m0
wV g

dry mass
mass of displaced water

(2.1)

is the density of water and Vg is the green wood volume.

Specific gravity in sawn-woods research is indefinable because the weight


of wood in a given volume changes with the shrinkage and swelling which
follows a change in moisture content. So, in wood testing a nominal specific
gravity is determined based on the volume of wood at the time of test and its
weight when oven-dried. Specific gravity varies considerably, not only between
species but also within individual pieces of the same species and its value must be
found by using standard deviation methods and plotting a Gaussian curve. Once
the specific gravity is known, the various strength tests on moisture content are
adjusted to take into account its effect upon these strength values.
The density or weight per unit volume of a piece of wood is a particularly
important property. Density is defined as:
m
[kg/m3]
(2.2)
V
41

Wood properties

where m is the mass of timber [kg] and V is its volume [m3].


at a moisture content
Density
shrinkage V, as:
m
V

m 0 1 0.01
V0 1 0.01 V

[%] is expressed, related to volumetric

1 0.01
1 0.01 V

(2.3)

Wood substance has a density of about 1500 kg/m3. Wood itself consists of
a mixture of wood substance and spaces, therefore the amount of wood substance
per unit volume decides the dry density, which can vary in common species from
about 300 kg/m3 to 800 kg/m3. If the spaces are filled with water this greatly
increases the density, so that green or living wood, which normally has a moisture
content of 60% to 200%, may have a density in excess of 1000 kg/m3.
The high strength to weight ratio is one of the principal advantages of wood
compared with competitive materials, although sometimes cheapness, ease of
working or low thermal conductivity may be the reason why wood has been
selected for a particular use. Wood is considered to have moderate density if its
dry density lies between about 360 and 500 kg/m3, so that woods below this range
are light woods and those above are heavy woods.
Romanian units normally refer to densities at the normal moisture content
of 15%. This standard value of moisture content is often applied when measuring
the physical and mechanical properties of wood. In other countries, the
standardised values of moisture content are different, refer to moisture content
values of 12% to 20%.

2.2.6. Moisture content


Moisture content, MC or , is the weight of water in the wood expressed as
a percentage of the weight of the oven-dry wood. It is determined by weighting
the sample prior to and immediately following drying and expressing the less of
weight as a percentage of the oven-dry weight.
original weight dry weight
100
dry weight

[%]

(2.4)

Drying is carried out in laboratory controlled conditions at a constant


temperature of 100 2 oC until the weight becomes constant. In its natural state,
wood contains a considerable amount of moisture held partly in the cell walls and
partly in the cell cavities. During seasoning, most of the water in the cell cavities
42

Timber structures - 2

is lost, leaving a condition known as the fibre-saturation point (FSP). Figure 2.2
shows how this phenomenon may be clarified by plotting the results of
compression tests on a particular species against varying percentages of moisture.
The graph shows that the fibre-saturation point occurs at around 25-30% and 25%
is generally accepted as being a norm in sawn lumber and timber strength
assessment. Between the fibre-saturation point and zero moisture content, wood
shrinks as it loses moisture and swells as it absorbs moisture. Above the fibresaturation point, there is no dimensional change with variation in moisture
content.
The trend of change of strength with change of moisture content is similar
for most strength properties but the magnitude of the change varies from one
property to another. The change in compression strength, for example, is more
than that for bending strength which in turn, is more than that found in the
modulus of elasticity. Decreases in moisture content below to the fibre-saturation
point are accompanied by increases in all strength properties.
Adjustments of strength values for percentages below the fibre-saturation
point are found by employing a mathematical equation derived by extensive
research and study. The obvious solution to all these problems is to use only wood
with low movement but this is not always realistic.
The moisture content of a piece of lumber obviously affects the cross
sectional dimensions: width and depth of a member, which are used to calculate
the section properties in structural design. Volume changes occur with moisture
content changes below the fibre-saturation point. If moisture is lost, wood shrinks;
if moisture is gained, wood swells.
[N/mm2]

Fibre-saturation point

MC
10

20

30

40

[%]

Figure 2.2 - Variation of strength versus moisture content


43

Wood properties

The volume changes that occur in wood when it loses moisture are greater
parallel to the annular ring than normal to the annular ring. The main concern is
the change in dimensions of the cross section of a piece of lumber and the radial
and tangential values are of primary interest. The volumetric shrinkage, V, is the
sum of the radial, tangential, and longitudinal shrinkage. The coefficient of
volumetric movement can be also considered to be equal to the numeric value of
density times 10-3. In other words, volumes of timber of a density equal to 400
kg/m3 swells 0.4% for each 1% increase in moisture content. The coefficient of
longitudinal movement, 0, is usually negligible in which case the coefficient of
transverse movement, 90, is equal to half of the coefficient of volumetric
movement. For most species, including spruce, pine, fir, larch, poplar and oak,
engineering values of 0 and 90 can be taken as 0 = 0.01 and 90 = 0.2. For dense
species like beech a 90 = 0.3 should be used, [23].
In Table 2.2 are given the fibre-saturation point values at room temperature
and typical radial and tangential shrinkage in drying from FSP (green) to ovendry of different woods.
Species
Fir tree
Larch
Pine
Spruce
Poplar

FSP [%]
30
28
21
27
31.5

Radial shrinkage [%]


4.5
4.5
5.1
3.8
2.1

Table 2.2
Tangential shrinkage [%]
7.9
9.1
7.5
7.8
6.1

The tangential movements may, for practical purposes, be taken as twice


the radial movements. For most engineering purposes, however, it is unnecessary
to differentiate between the two transverse directions, and transverse movement is
often taken as the average value. Changes in dimensions tend to be linear with
moisture in the range of 5 to 20% moisture content. In this range movements can
be calculated from:
h2

h1 1

100

(2.5)

where: - h1 and h2 are the dimensions at moisture 1 and 2;


- is the coefficient of swelling (positive) or shrinkage (negative).
These non-uniform volume changes can result in seasoning defects known
as checks, which are radial cracks.
Fortunately for the designer, it is not necessary to compute section
properties based on a consideration of the initial moisture content, final moisture
content, and the resulting shrinkage (or swelling) that occurs in the member.
44

Timber structures - 2

Lumber grading practices have established the dry size of lumber as the basis for
structural calculations. This means that only one set of cross-sectional properties
needs to be considered in design. This is made possible by fabricating lumber to
different cross-sectional dimensions based on the moisture content of the wood at
the time of fabrication. Therefore, lumber, which is fabricated from green wood,
will be somewhat larger at the time of fabrication. However, when this wood
reaches the dry moisture content condition, the cross-sectional dimensions will
closely coincide with those for lumber fabricated in the dry condition. Lumber
with 20% and over in moisture content is defined as unseasoned or green lumber.
The moisture content of structural lumber in service is much lower. The
average moisture content that lumber assumes in service is known as the
equilibrium moisture content (EMC). Depending on the atmospheric conditions,
the EMC of structural framing lumber in a covered structure (dry conditions) will
range somewhere between 7% and 14%. In most cases, the MC (moisture content)
at the time of construction will be higher than the EMC of a building (perhaps 2
times higher).
The drying of lumber in order to increase its structural properties is known
as seasoning. As noted, the MC of lumber in a building typically decreases after
construction until the EMC is reached. Although this drying in service can be
called seasoning, the term seasoning often refers to a controlled drying process.
Air-drying or kiln drying can perform controlled drying and both increase the cost
of lumber.
In order to stabilise the moisture content the only obvious solution is to
enclose the wood within a protective film. Paint and varnish coatings will act in
this way, if they completely cover the wood and they are not damaged in any way.
Unfortunately, they are unable to prevent moisture content changes resulting from
slow seasonal changes in atmospheric relative humidity. This failure is known as
preferential wetting and is responsible for blistering and peeling in paint wood,
and the loss of transparency in varnishes.

2.3. MECHANICAL PROPERTIES


The aim of the mechanical properties discussion is to relate the
characteristics of wood directly to the needs of wood structure design. The
mechanical properties become the bases for the allowable strengths or design
stresses given by design codes and specifications.
Traditionally, many of the mechanical properties of wood have been
determined on tests of small pieces that are termed clear and straight - grained
because they did not contain defects such as knots, cross grain, checks, and splits.
They will contain, of course, growth rings that occur with regular or consistent
patterns within the pieces. In mechanics the wood specimens containing growth
rings are generally considered homogeneous. Non-homogeneity in wood members
45

Wood properties

is associated with the more random presence of characteristics of wood that were
discussed. Since much wood testing is carried out on the small, clear, straightgrained specimens, it is often important in design application to distinguish
between results of tests based on small, clear, straight-grain specimens and the
performance of full-sized or fabrication-modified structural elements. The former
often indicate important general relationships, [2].
The strength and stiffness properties of most interest in structural design
are:
compressive strength parallel to the grain;
compressive strength perpendicular to the grain;
tensile strength parallel to the grain;
bending strength;
shear strength;
modulus of elasticity parallel to the grain;
shear modulus.
These strength and stiffness properties are determined by a series of
standardised tests.

2.3.1. Stiffness properties


A result of the orthotropy of wood is the three moduli of elasticity, three
shear moduli and six Poisson's ratios.
The modulus of elasticity, also called Youngs modulus, usually used in the
design process is taken as a longitudinal modulus, EL. Data for ER and ET are not
extensive and usually they are not presented as allowable properties for species.
However, where a transverse modulus, ET (or E ), is essential in design, an
approximation often used is 0.06 times the longitudinal value.
The values of EL and the ratios with other constants vary within and
between species and with moisture content and specific gravity. A typical
coefficient of variation for EL in clear wood of a single species is 22 %.
GLR, GLT and GRT denote the three moduli of rigidity, or shear moduli, in
the (LR), (LT) and (RT) planes respectively.
The six Poisson's ratios are denoted by LR, RL, LT, TL, RT and TR.
The first letter of the subscript refers to the direction of applied stress and the
second letter refers to the direction of lateral deformation. For example: LR is the
Poisson's ratio for deformation along the radial axis caused by stress along the
longitudinal axis and its values are between 0.30 and 0.45 depending on the
toughness of wood.
The Romanian codes STAS 856-71 and NP005-96 present the design
values of: elasticity modulus in longitudinal (parallel) direction, EL, elasticity
46

Timber structures - 2

modulus in transverse (perpendicular) direction, ET, and shear modulus GRT for
softwood and hardwood. These values are evaluated for a wood moisture content
of 15%. Table 2.3 presents some of these values.

softwood
hardwood

Modulus of elasticity [N/mm2]


EL (E//)
ET (E )
10,000 11,300
300
11,500 14,300
600

Table 2.3
Shear modulus [N/mm2]
GRT
500
1000

2.3.2. Strength properties


The strength properties of wood are tabulated in several ways. The first is a
compilation of the properties of clear, straight - grained specimens. Research
laboratories have studied the properties of wood for structural purpose through
tests of small pieces. The properties of clear, straight - grained wood have two
major purposes. They serve as a source for derivation of allowable properties and
provide a basic reference point for making species selections for a specific end use
where ranking of available species is required. Often the small specimen
properties are presented only as averages for a species or for a particular sample
subset of the species. If these clear wood properties are to be used for engineering
purposes, the variability inherent in the property should be considered as well as
the influence that growth and environmental factors have on the property for clear,
straight - grained wood.
Strength properties are also tabulated as allowable properties for design. A
major advantage of using allowable properties for engineering applications is that
considerations of variability and anticipated end use have been made by standard
techniques. Techniques used often are documented as: STAS 856 - 71 "Wood
Constructions. Design rules"; ASTM Standard D2555 "Establishing Clear-Wood
Strength Values"; D245 "Establishing Structural Grades and Related Allowable
Properties for Visually Grader Lumber"; D2915 "Evaluating Allowable Properties
for Grades of Structural Lumber"; AHA 1S 1-71" Interim Industry Standard
Developed by the American Hardboard Association".
The more common mechanical properties are normally carried through the
derivation to give allowable properties. These usually include the bending
strength, compression strength perpendicular and parallel to the grain, tensile
strength parallel to the grain and shear strength. Most allowable properties in
bending, compression, tensile and shear are presented in the Romanian codes
STAS 856-71 and NP005 96.
Strength properties less commonly measured on a wide variety of species
and thus not often presented for design use include torsion, toughness, creep,
47

Wood properties

rolling shear, fatigue resistance, sound transmission and internal friction. For
design application, efforts are often made to relate the less common properties to
those more commonly measured.
Torsion shear strength, for design purpose, is usually taken as equivalent
to the shear strength parallel to the grain.
Toughness represents the energy required to cause a complete failure under
very rapid loading.
Fatigue resistance of wood is sometimes an important consideration in
design. Because wood is fibrous it tends to be less sensitive to repeated loads than
some crystalline materials. Studies on the fatigue character of wood have not been
extensive and trend to be limited to several modes of loading, several range ratios
of minimum to maximum stress, and several moisture content conditions.
Rolling shear is described as the shear strength of wood where the shearing
face is the longitudinal - tangential plane and perpendicular to the grain.
In Table 1, the Romanian STAS 856-71 presents the common allowable
strengths of different wood species for permanent and short-term use timber
constructions. These values must be adjusted by coefficients in the cases of
different environmental and use conditions (clauses 4.1.2 4.1.12, STAS 85671). Summaries of the allowable strengths are presented here in Table 2.4.
The Romanian code NP005-96 presents the characteristic strengths
according to the wood species and quality classes in table 2.3. Table 2.5, shows
here these characteristic strengths for traditionally wood species.
Note: A characteristic value is defined as a population lowers 5-percentile value,
which must be evaluated experimentally.
Type of load
1
2
3
4
5
6
7
8

48

Bending allowable strength


Compression allowable strength parallel to the grain
Compression allowable strength perpendicular to the
grain
Tension allowable strength parallel to the grain
Bearing allowable strength parallel to the grain
Bearing allowable strength perpendicular to the grain
Shear allowable strength parallel to the grain
Shear allowable strength perpendicular to the grain

Table 2.4
Allowable strength [N/mm2]
Symbol Softwood Hardwood
10
11 - 13
ai
10
11 - 13
ac//
1.5
2.4 3.0
ac
at
as//
as
af//
af

7 8.5
10
1.5
2
4.5

8.5 - 10
11 - 13
2.4 3.0
2.6 - 3.2
6 7.5

Timber structures - 2

Ri

Table 2.5
Characteristic strength [N/mm2]
Softwood
Hardwood
I
II
III
I
II
III
24.0
16.8
9.6
40.0
28.0
16.0

Rc//

15.0

12.0

4.5

19.8

15.8

5.9

Rc

3.3

3.0

10.4

9.4

Rt

14.4

8.6

4.3

22.5

13.5

6.8

Rf//

3.0

2.7

6.4

5.7

Rf

12.0

10.8

24.0

21.6

Type of load
Symbol
1
2
3

4
5
6

Bending characteristic
strength
Compression characteristic
strength parallel to the grain
Compression characteristic
strength perpendicular to
the grain
Tension characteristic
strength parallel to the grain
Shear characteristic
strength parallel to the grain
Shear characteristic
strength perpendicular to
the grain

2.4. STRENGTH CLASSES


The European Common Market will lead to a more varied timber supply in
most European Union (EU) and European Free Trade Agreement (EFTA)
countries, with a correspondingly larger number of grades and characteristic
values. To keep the specification process of timber simple and to avoid confusion,
a strength class system is being introduced. A strength class system has been
established in EN 338 and the Romanian code SR-EN 338.
The introduction of strength classes is advantageous to both timber user and
timber supplier. The designer does not need to acquaint himself with a multitude
of different grades and related characteristic values, no matter in which European
country his/her project will be built. Instead, he can simply choose the strength
class suitable for his project from a concise table, similar to those used for other
structural materials. The timber producer has the advantage that he can achieve
higher prices for his timber since the better the grading process applied, the higher
the strength classes to which his timber is allocated. Grading machines can be
used to grade the timber directly into strength classes and also into classes, which
could not be achieved by visual grading. The timber supplier has the advantage
that he can select the most economic source for a specific grade.
The strength class system established in SR-EN 338 Structural timber
Strength classes is shown in Tables 2.6.a and 2.6.b, which give the
characteristics strength and stiffness properties and density values for each
strength class. Table 2.6.a presents the strength classes and characteristic values
for 9 classes of coniferous species and poplar whereas Table 2.6.b shows the same
values for 6 classes of deciduous species. It ranges from the weakest grade of
softwood, C14, to the highest grade of hardwood, D70, currently used in Europe.
49

Wood properties

C14

C16

C18

fm,k
ft,0,k
ft,90,k
fc,0,k
fc,90,k
fv,k

14
8
0.3
16
4.3
1.7

16
10
0.3
17
4.6
1.8

18
11
0.3
18
4.8
2.0

E0,mean
E0,05
E90,mean
Gmean

7
4.7
0.23
0.44

8
5.4
0.27
0.50

9
6.0
0.30
0.56

290

310

320

D30

D35

fm,k
ft,0,k
ft,90,k
fc,0,k
fc,90,k
fv,k

30
18
0.6
23
8.0
3.0

E0,mean
E0,05
E90,mean
Gmean

10
8.0
0.64
0.60

530

D40
[N/mm2]
35
40
21
24
0.6
0.6
25
26
8.4
8.8
3.4
3.8
[kN/mm2]
10
11
8.7
9.4
0.69
0.75
0.65
0.70
[kg/m3]
560
590

C22
C24
[N/mm2]
22
24
13
14
0.3
0.4
20
21
5.1
5.3
2.4
2.5
[kN/mm2]
10
11
6.7
7.4
0.33
0.37
0.63
0.69
[kg/m3]
340
350

C27

C30

Table 2.6.a
C35
C40

27
16
0.4
22
5.6
2.8

30
18
0.4
23
5.7
3.0

35
21
0.4
25
6.0
3.4

40
24
0.4
26
6.3
3.8

12
8.0
0.40
0.75

12
8.0
0.40
0.75

13
8.7
0.43
0.81

14
9.4
0.47
0.88

380

400

420

D50

370
Table 2.6.b
D60
D70

50
30
0.6
29
9.7
4.6

60
36
0.7
32
10.5
5.3

70
42
0.9
34
13.5
6.0

14
11.8
0.93
0.88

17
14.3
1.13
1.06

20
16.8
1.33
1.25

650

700

900

m = bending;
t = tension;
c = compression;
v = shear;
f = strength
k = characteristic;
0 = parallel to the
grain;
90 = perpendicular to
the grain.

Experimental data show that all-important characteristic strength and


stiffness properties can be approximated from either bending strength, modulus of
elasticity or density. These relationships, according to EC5, are:
(2.6)
(2.7)
f t , 0 , k 0.6 f m, k
f c , 0 , k 5 f m0,.k45
f v , k 0.2 f m0,.k8
(2.8)
(2.9)
f c , 90, k 0.015 k
(2.11)
(2.10)
E 0.05 0.67 E 0, mean
f t , 90, k 0.001 k
E 0, mean
E 0, mean
(2.12)
(2.13)
G mean
E 90, mean
16
30
Due to the relationships between strength, stiffness and density shown
above a species/source/grade combination can be assigned to a specific strength
class based on the characteristic values of bending strength, modulus of elasticity
and density.
50

Timber structures - 2

2.5. INFLUENCE OF VARIOUS FACTORS ON WOOD


PROPERTIES
2.5.1. Density
Much of the difference in properties between and within species can be
attributed to differences in density. The specific gravity of the wood substance is
essentially constant at about 1.50, so the specific gravity of a piece of wood is
effectively a measure of the proportion of the volume occupied by wood
substance. Since it is the wood substance that imparts strength to a piece of wood,
the greater the proportion of wood substance, the higher the mechanical properties
may be expected to be. The relation between mechanical properties and specific
gravity has the form:
S

KG n

(2.14)

where: - S = the value of any particular mechanical property


- G = specific gravity
- K, n = constants depending on the particular property being considered.
For differences between species, the values of n range from 1.00 to 2.25,
and for within-species variation, they range from 1.25 to 2.50.
As long as the wood is clear and straight grained, specific gravity is a useful
predictor of strength. When the wood is not clear or straight grained and contains
other growth characteristics or manufacturing imperfections, the influence of
specific gravity may be diminished, depending upon the property and the severity
of the characteristics in the piece.

2.5.2. Moisture content


Changes in moisture content below the fibre saturation point result in
changes in the mechanical properties, with higher properties at the lower moisture
contents, except that properties related to impact (toughness, work to maximum
load in static bending, and height of drop in impact bending) may remain the same
or may actually decrease at the lower moisture contents.
Mechanical properties increase with decrease in moisture content. Most
clear wood mechanical properties obey the following relation in the vicinity of
20oC:
MC1 MC 2
MG MC 2
P
PMC1 PMC 2 MC 2
PMG
where: - PMC1 = property at moisture content MC1;

(2.15)

51

Wood properties

- PMC2 = property at moisture content MC2;


- PMG = value of property for all moisture contents greater than moisture
content MG (slightly below fibre saturation point), at which
property changes due to drying are first observed.
The formula assumes small, clear specimens and a drying process in which
no deterioration or product degrades occur.
Since mechanical property values are commonly given for the green
condition and at a moisture content of 15%, MC2 may be taken as 15 and:
MC1 15
MG 15

P15
PMG
Values of MG for a few species are given in Table 2.7.
PMC1

(2.16)

P15

Table 2.7
Species
Ash, white
Birch, yellow
Larch
Pine
Spruce

MG [%]
24
27
28
21
27

Values for the effect of moisture on the mechanical properties of clear


wood properties are given in Table 2.8. For practical purposes a linear relationship
between moisture content and properties may be assumed for 8% < < 20%.
Table 2.8
Property
Compression strength parallel to the grain
Compression strength perpendicular to the grain
Bending strength parallel to the grain
Tension strength parallel to the grain
Tension strength perpendicular to the grain
Shear strength parallel to the grain
Modulus of elasticity parallel to the grain

Change [%]
5
5
4
2.5
2
3
1.5

In timber design the influence of moisture is taken into consideration by


assigning timber structures to service classes. The European code EC5 and the
Romanian code NP005 96 define this modification factor, mui. Code NP005
96 in table 2.5 gives the following values (subscript i defines the load type):
- 1.00 for all types of loads and the first service class of the timber
construction;
- 0.90 for all types of loads and the second service class of the timber
construction;
- 0.70 0.90 for the third service class of the timber construction and
different loads.
52

Timber structures - 2

2.5.3. Knots
The influence of a knot on the mechanical properties of a product varies
depending upon the size, location, and type of stress that is applied to the member.
Knots decrease the mechanical properties because the knot displaces clear
wood, because of stress concentrations that occur due to the discontinuities of
wood fibres, and because of the deviation of the fibre growth around the knot. In
general, the strength-reducing effect of knots depends on the portion of the cross
section that they occupy.
A knot can have a more serious effect on the tension side of a member
loaded in bending than it does in compressive side.
In long columns knots are important because they affect stiffness as well as
strength. In round timbers, such as poles and piles, knots have less affect because
the material has not been sawn.
When loads are applied perpendicular to grain, knots actually increase both
the hardness and the strength, but they may be objectionable in such applications
because they result in uneven wear or stress concentrations.

2.5.4. Fibre and ring orientation


The directions of important stresses in wood member may not coincide
with orthotropic axes of fibre orientation.
The allowable properties of wood products are given by standard
procedures. The influence of fibre direction on mechanical properties can be
approximated by Hankinson's formula, [29]. This equation (2.17) relates the
property at any angle of the fibre direction to that property transverse and parallel
to the grain.
N

P sin n

PQ
Q cos n

(2.17)

where: - N = the property at an angle ;


- = the angle between property direction and direction
parallel to the grain;
- Q = the property across the grain;
- P = the property parallel to the grain;
- n = empirically determined constant.
Empirically derived constants are tabulated for the different properties of
most interest. The values of n and associated Q/P ratios have been taken from
available literature and are presented in Table 2.9.
53

Wood properties

Table 2.9
Property

Q/P

Tensile strength

1.5 - 2

0.04 - 0.4

Compressive strength

2 - 2.5

0.03 - 0.4

Bending strength

1.5 - 2

0.04 - 0.1

Toughness

1.5 - 2

0.06 - 0.1

0.04 - 0.12

Modulus of Elasticity

2.5.5. Temperature
Temperature can be an important environmental condition in specific
applications. There are both immediate and permanent effects of temperature. An
immediate effect means that the increase or decrease in property associated with
temperature change is not retained once the member returns to its original
temperature. On the other hand, permanent effects can take place if, for example,
heating takes places over a long period of time. As a practical matter, the length of
time heat that can be tolerated before a permanent effects takes place is a function
of temperature. Heating under conditions of high humidity, such as steam or hot
water is more serious than under dry conditions. Thus a beam near a cooking vatemitting steam at high temperature would suffer more than would a similar beam
subjected to dry heat.
Property
[percent of value at 20oC]

200

100

Temperature
-200

-100

+20

+100

+200

[oC]

Figure 2.3 Immediate effect of temperature on strength properties


54

Timber structures - 2

An example of an immediate increase in strength as wood is cooled below


20 C is shown in fig.2.3. The figure illustrates that above 20oC the strength of
wood decreases quite linearly up to approximately 200oC. The two bands illustrate
the influence of moisture content. The width of the bands represents variability in
experimental results reported in the literature, [10].
Effects of temperature on other properties of wood in different heating
media are available in the literature. Extremes of temperature are obviously
reached when a wood member is subjected to fire. The behaviour of wood when
exposed to fire is discussed in the dedicated chapter.
o

2.5.6. Duration of load


Apart from the moisture content, the duration of the load significantly
influences the strength and deformations of timber and timber structures. As load
duration increases, the strength of timber decreases.
When loaded for the first time, a wooden member deforms elastically. If the
load is maintained, an additional time-dependent deformation termed creep
occurs. Even at very low stresses creep takes place and can continue for several
years. Unloading the member, of course, results in immediate and complete
disappearance of the original elastic deformation and delayed recovery of
approximately one-half of the creep. If, instead of controlling the load, a constant
deformation is imposed and maintained on a wood member, the initial stress
relaxes at a decreasing rate to about 60% of its original value within a few months.
This reduction of stress is termed relaxation. In a sense this is the inverse or
mirror image of the creep.
In addition to the creep/relaxation characteristic of wood, there is a duration
of stress or time - dependent effect on the strength of wood. The load required to
produce failure after a 10-year period of loading, for example, will be
approximately 60% of the load that might have normally been required in a 5minute structural test. On the other hand, wood on impact will support a higher
load than that predicted by the 5-minute test, [25].
Fatigue should be considered in wood design when repetitions of design
stress are expected to be more than 100,000 cycles during the normal life of a
structure. There are several technical considerations involved, including the
fatigue life, the range ratio, and the type of loading. The literature is not extensive
on the fatigue testing of wood. It does suggest, however, that typically a clear,
straight-grained bending specimen with a fatigue life of 2 million cycles will have
a strength of approximately 60% of what was anticipated under static conditions.
Knots and slope grain appear to be additive. After 2 million cycles specimens with
a slope of grain of 1:12 and small knots have shown a strength of about 30 percent
of the clear, straight-grained static value, [10].
55

Wood properties

In the design process this load duration effect on the wood properties is
taken into consideration as a factor, mdi, generally called working condition
coefficient or modification factor. It is defined according to the duration of load
and the wood species. Table 2.10 presents values of these coefficients according
to NP00596 (table 2.6).
Table 2.10
Type of load
Static bending
Shear

Compression

Tension

Elasticity modulus

Load duration class


Permanent load
Long term variable
load
Short term variable
load
Permanent load
Long term variable
load
Short term variable
load
Permanent load
Long term variable
load
Short term variable
load
Permanent load
Long term variable
load
Short term variable
load

mdi

Symbol

mdi
mdf

softwood
0.55
0.65

hardwood
0.60
0.70

1.00

1.00

0.80
0.85

0.85
0.90

1.00

1.00

0.90
0.95

0.95
1.00

1.00

1.00

1.00
1.00

1.00
1.00

1.00

1.00

mdc

mdt

mdE

2.5.7. Chemicals and decay


Chemicals may degrade wood, the degree of degradation being reflected in
varying degrees of loss in mechanical properties. The effect of chemicals on
mechanical properties is highly dependent upon the specific type of chemicals.
Some typical strength retentions after chemical exposure as a percentage of watersoaked control, [14], are given in Table 2.11.
Concentration
[%]

Nitric
[%]

2
6

80
60

2
6

44
13

56

Sulphuric
[%]
Fir tree
92
89
Oak tree
80
60

Acetic
[%]

Table 2.11
Caustic
[%]

90
88

22
0

101
101

20
0

Timber structures - 2

Fire retardants based on salts are impregnated into wood in high


concentrations and have a substantial effect on its strength. For example, the
allowable unit stresses for solid timber treated with waterborne salts must be
decreased by 10%.
Wood-destroying fungi seriously reduce strength. Some fungi cause easily
recognised pitting, others give only a slight discoloration in the early stages. No
method is known for estimating the amount of reduction in strength from the
appearance of decayed wood.
One measure of the progress of decay is the amount of weight loss as a
result of fungal attack. A summary of the estimated values for strength loss in
softwoods and hardwoods at early stages of decay is given in Table 2.12, [8].

Toughness
Impact bending

General bending strength

Work to maximum load


Modulus of rupture

Modulus of elasticity

Compression, perpendicular (radial)

Compression, parallel
Tension, parallel

Shear, parallel
Hardness

Table 2.12
Estimated values for strength loss
Softwood
Hardwood
Approximate weight loss Approximate weight loss
[%]
[%]
1 2 4 6 8 10 1 2 4 6 8 10
57
75
36
60
20 20 25 62
6 31 60
8 70
78 85
80
38 50 55 72
27 50 70
89 92
13
36
5
16
34
27
54 69 75
13
61
70
32 49 61
50
4
66
55
18 25
48
6
16
48
66
19
24 35
60
10
25
10
25
45
23
60 50
56
82
40
2
6
15 20
7
21

Insects may destroy most of a piece of wood, frequently without external


evidence of the damage. There is no known method for estimating the strength
reduction from insect damage. Therefore, when strength is an important
consideration the safe procedure is to discard pieces that contain decay or insect
damages.
57

Wood properties

Try to
answer
these

Questions for you

1. Name the wood properties of interest for designer.


2. List the wood physical properties and explain one of them.
3. Define the following terms: moisture content, fibre-saturation point and
equilibrium moisture content.
4. List the wood mechanical properties and give some detail for each of them
5. List the main three factors, which influence the wood properties and explain
one of them in detail.
6. A fir beam measures 15 cm width and 20 cm depth, when its MC is 10%.
What will its width and depth be when MC is raised to 35%? What are they
when MC is raised from 10% to 19%? Draw the diagrams of the width and
depth variation versus MC increase from 0% to 100%.

58

Timber structures - 3

3.1. DESIGN METHODS


Structures shall be designed and constructed safely and economically as
well as properly for the purposes. The objective of design is to assure that the
structure and the members possess adequate safety against all loads during both
service and construction stages, and to provide their functions efficiently for
ordinary services. In addition, it shall be considered that structures possess
sufficient durability, and match the environment during their lifetime.
There are two design methods, which can be used in wood engineering
structures design:
- allowable strengths design method (ASDM);
- limit states design method (LSDM).

3.1.1. Allowable strengths method


3.1.1.1 Design philosophy
The allowable strength design philosophy is a method in which design
stresses are derived on a statistical basis and deformations are also limited. This
method is based on the Bernoulli's hypothesis (the cross-sections are plane and
perpendicularly on the axes of the element before and after its loading
deformation) and the Hook's rule (the relationship between stress and strain is
given by linear analysis. The stress and the strain are available for elastic
behaviour of element only).
Allowable strength (allowable stress or permissible stress) is a unique
value which results from laboratory tests. This value represents the maximum
stress, which can appear in the wood element for a certain load.
59

Design methods. Actions

The allowable strengths are based on the failure strengths of the small, clear
specimens and a drying process in which no products deterioration or degradation
occurs. Thus the allowable strengths for each specimen and each load (action) are
obtained from the results of standard tests on a small clear specimen, by dividing
the statistical medium value of failure strength by the unique appropriate safety
factor for material and manufacture conditions.
The failure strengths are noted and defined as:
-

r
r

= failure strength;
= failure shear strength.

The unique safety factor is noted c, therefore:


r
a

where: -

a
a

and

r
a

c
c
= allowable strength;
= allowable shear strength.

(3.1)

The value of the unique safety factors are based on the analysis of:
- wood defects;
- moisture content;
- fibre and ring orientation;
- the difference between the performance of full-sized structural
elements and the results of tests based on small, clear, straightgrain specimens;
- stress concentrations that occur due to the discontinuities of wood
fibres;
- the possibility to increase the loads more than they were
considered during design;
- load sharing;
- the design errors.
The maximum values of actual stresses or actual displacements are usually
noted: max,ef, max,ef and fmax,ef.
In the allowable strength design philosophy, the maximum effective
stresses and displacements should satisfy the following condition:
max,ef

(a)

(b)
f max,ef f a
(c)
in which fa is the allowable displacement.
max,ef

60

(3.2)

Timber structures - 3

3.1.2. Limit states method


3.1.2.1. Design philosophy
The basic stresses in allowable strength methods were determined by
carrying out short-term loading tests on small timber specimens free from all
defects. The data were used to estimate the minimum strength, which was taken as
the value below which not more than 1% of the test results fell. These strengths
were multiplied by a reduction factor to give basic stresses (or allowable
strengths). Research studies had shown the need for a review of the stress values
and safety factor. The new approach for assessing the strength of timber moved
somewhat in line with limit states design philosophy.
Limit states are states beyond which the structure no longer satisfies the
design performance requirements. Limit states are classified into:
- ultimate limit states;
- serviceability limit states.
Ultimate limit states are those associated with collapse, or with other forms
of structural failure which may endanger the safety of people.
In Table 3.1 some examples of the ultimate limit states are given.
Table 3.1
Ultimate limit states
Rupture of section
Stability
Displacement
Deformation

Mechanism

Condition
Rupture of critical section of structural members.
Loss of stability of the whole or a part of the structure as a
rigid body by overturning or other motions.
Loss of load carrying capacity of the structure due to large
displacement.
Loss of load carrying capacity of the structure due to
excessive deformation by plastic deformation, creep,
cracking and differential settlement.
Transformation of the structure into a mechanism.

Ultimate limit states that may require consideration include:


- loss of equilibrium of the structure or any part of it, considered as
a rigid body;
- failure by excessive deformation, rupture, or loss of stability of
the structure or any part of it, including supports and foundations.
Serviceability limit states correspond to states beyond which specified
service requirements are no longer met.
The serviceability limit states are associated with normal use or durability
of the structure. Table 3.2 shows examples.
61

Design methods. Actions

Serviceability limit states


Cracking
Deformation

Local damage
Vibration

Table 3.2
Conditions
Impairment of appearance, durability, or water
and air tightness of the structure.
Excessive deformation, which does not impair
stability and equilibrium, but is not suitable for
normal use.
Local damage which prevents normal use of
the structure.
Excessive vibration which is not suitable for
normal use, or produces uneasiness.

Serviceability limit states that may require consideration include:


- deformations or deflections, which affect the appearance or
effective use of the structure (including the malfunction of
machines or services) or cause damage to finishes or nonstructural elements;
- vibrations, which cause discomfort to people, damage to the
building or its contents, or which limit its functional
effectiveness;
- cracking of the wood which is likely to affect appearance,
durability or water tightness adversely;
- damaging of wood in the presence of excessive compression,
which is likely to lead to loss of durability.
The structural design of timber members according to the limit states
method is based on permissible stress design philosophy in which design stresses
are derived on a statistical basis and deformations are also limited. Elastic theory
is used to analyse structures under various loading conditions to give the worst
design case. The safety verification is based on the partial coefficient method.
The main parameters are the actions, the material properties and the
geometrical data. Normally, these parameters are stochastic variables with
distribution functions as shown in principle in fig. 3.1 for the action effects (S) and
the corresponding resistance (R), [26].
The distributions have the mean values Smean and Rmean and they can be
assigned characteristic values Sk and Rk defined as fractiles in the distribution. For
actions an upper fractile is normally used and for resistance a lower fractile or the
mean value may be appropriate. The purpose of the design is to get a low
probability of failure, i.e. a low probability of getting action values higher than the
resistances. This, in the partial coefficient method, is achieved by using design
values found by multiplying the characteristic actions and dividing the
characteristic strength parameters respectively, by partial safety coefficients.
62

Timber structures - 3

0
Smean

Sk

Rk

Rmean

Figure 3.1 Statistical distribution (idealised) for action effects (S) and
resistance (R)
In all relevant design situations, it must be verified that the limit states are
not reached when design values for actions, material properties and geometrical
data are used in the design models. In particular it must be verified that:
- the effects of design actions do not exceed the design resistance at
the ultimate limit states;
- the effect of design actions does not exceed the performance
criteria for the serviceability limit states.
In symbolic form, for ultimate states related to rupture, it must be verified
that:
S d Rd
(3.3)
For ultimate states related to static equilibrium or to gross displacement of
the structure as a rigid body, the corresponding expression is:
S d ,dst

S d ,stb

(3.4)

For serviceability limit state it shall be verified that:

Sd

Cd

(3.5)

where: - Sd = the design value of the effects of the actions such as axial force,
moment or a vector of several forces and moments, displacement or
acceleration;
- Rd = the corresponding design resistance;
- Sd,dst = the design value of the effect of destabilising actions;
- Sd,stb = the design value of the effect of stabilising actions;
- Cd = a prescribed value.
63

Design methods. Actions

3.1.2.2. Design values


Design values of actions
The design actions may be different for the different limit states. Firstly, the
possible load cases are identified, i.e. compatible load arrangements, sets of
deformations and imperfections. A load arrangement identifies the position,
magnitude and direction of an action. Secondly, the actions are combined
according to the following symbolic expressions:
- for ultimate limit states
- fundamental load combinations
G , j Gk , j
i
Q ,i Qk ,i
- special load combinations
Gk , j
Ak ,l
i Q k ,i

(3.6.a)
(3.6.b)

- for serviceability limit states


Gk , j

Qk , i

(3.7)

where: - = the partial factors or load factors;


- = the combination factors;
- G = the permanent actions;
- Q = the variable actions;
- A = the accidental actions.
The representative values multiplied by - values are called design actions.
Finally, the effects (S) of actions for example internal forces and
moments, stresses, strains and displacements are determined from the design
values of the actions (F), geometrical data (a), and where relevant, material
properties (X):
Sd

S Fd ,1 , Fd ,2 ,.......a d ,1 , a d ,2 ,....... X d ,...

(3.8)

Design values of resistances


The design value Xd of a material property with the characteristic value Xk
is defined as:
X
X d k mod k
(3.9)
i

where: - I = partial safety factor for the property of the material;


- kmod = modification factor taking into account the effect on the strength
64

Timber structures - 3

parameters of the duration of the actions and the moisture content;


k mod

m ui m di

(3.10)

- mui and mdi were defined in chapter 2 (paragraphs 2.5.2 and 2.5.6);
- Xk = the characteristic value of the material property.
It is generally assumed that the relationship between the resistance, R, and
the strength parameters, f, the stiffness parameters, E, and the geometrical data, a
is known. If this is the case, design values should be used to determine the design
resistance:
Rd

R f 1,d , f 2 ,d ,....E1,d , E 2 ,d ,...a1,d , a 2 ,d ,...

(3.11)

The design value Rd can also be determined directly from characteristic


values, Rk determined from tests:
Rd

k mod

Rk

(3.12)

3.2. ACTIONS
For the intended construction work, the designer is first faced with the
conceptual design of the structural system. This stage will consider the type of
structure and the construction material to be used. The structural design then starts
with an analysis of the actions that may be applied to the chosen structure.
Account should be taken of direct actions that are the applied external forces as
well as the indirect actions that result from imposed deformations.
An action is:
a force (load) applied to the structure (direct action);
an imposed deformation (indirect action), for example,
temperature effects or settlement.
In addition to the previous classifications, differentiation of the actions has
to be considered according to the variation of their magnitude in space and with
time:
a) - by their variation in time:
- permanent actions (G or P), e.g. self-weight of structures,
fittings, ancillaries and fixed equipment;
- variable actions (Q or V) which are also classified in:
- long term variable actions, e.g. snow load;
- short term variable actions, e.g. wind load.
65

Design methods. Actions

- accidental actions (A or E), e.g. explosion or earthquake.


b) - by their spatial variation:
- fixed actions, e.g. self-weight;
- free actions, which result in different arrangements of actions,
e.g. movable imposed loads, wind loads, snow loads.
Indirect actions are either permanent (e.g. settlement of support) or variable
(e.g. temperature) and are treated accordingly. Supplementary classifications
relating to the response of the structure are given in the relevant clauses.
The following loads are of primary concern to a building designer:
1. Gravity loads
- dead load;
- live load;
- snow load.
2. Lateral loads
- wind load;
- seismic load;
- special loads and load effects (temperature influence variations,
structural foundation settlements, impact, and blast).
Characteristic values of actions are specified:
- in codes for actions or other relevant loading codes. The
Romanian standards and codes for actions are: STAS 10101/0-75,
STAS 10101/21-92, STAS 10101/20-90, STAS 10101/0A-77,
STAS 10101/0B-87, STAS 10101/23-75(78), STAS 10101/1-78,
STAS 10101/2-75, STAS 10101/23-75, P100/92 etc.
- by the client, or the designer in discussion with the client,
provided that minimum provisions, specified in the relevant codes
or by the competent authority, are observed.
Normally, a design specification does not prescribe the magnitudes of the
loads that are to be used as the basic input to the structural analysis. It is the role of
the specification to detail the methods and criteria to be used in design process.
The specification therefore reflects the requirements that must be satisfied by the
structure in order that it will have a response that allows it to achieve the needed
performance. Loads, on the other hand, are governed by the usage or type of
occupancy of the building, which in turn is dictated by the applicable local,
regional, and natural laws that are more commonly known as building codes.
66

Timber structures - 3

The building code loads have traditionally been given as nominal values,
determined on the basis of material properties (e.g. dead load) or load surveys
(e.g. live load and snow load). To be reasonably certain that the loads are not
exceeded in a given structure, the code values have tended to be higher than the
loads on a random structure at an arbitrary point in time. This may, in fact, be one
of the reasons why excessive gravity loads are rarely the obvious cause of
structural failures. Be that as it may, the fact of the matter is that all of the various
types of structural loads exhibit random variations that are functions of time, and
the manner of variation also depends on the type of load. Rather than dealing with
nominal loads that appear to be deterministic in nature a realistic design procedure
should take load variability into account along with that of the strength, in order
that adequate structural safety can be achieved through rational means.
Since the random variation of loads is a function of time as well as of a
number of other factors, the modelling, strictly speaking, should take this into
account by using stochastic analyses to reflect the time and space
interdependence. For most design situations the code will specify the magnitude
of the loads as if they were static. Their time and space variation are covered
through the use of the maximum load occurring over a certain reference (return)
period, and its statistic.
The geographical location of the structure plays an important role for
certain loads. It is particularly applicable to snow, wind, and seismic loads.
The loads on the structure are normally assumed to be independent of the
type of structure and structural materials, with the exception of dead loads. The
response of a building, however, will be different for different materials.,
depending on the type of load.
The size of a structure (height, floor area) has a significant impact on the
magnitudes of most loads. All loads are influenced by the increasing height of a
multi-story building, for example.
All these aspects are covered by the use of load factor, which is given in
codes and it multiplies the nominal load value for giving the maximum its value.
There are many types of loads that may act on a building structure at one
time or other. After the estimation of the actions, the design requires the structural
analysis of the action effects. This stage involves the selection of realistic load
arrangements for which the structure or the structural components are to be
designed. Then the design values result from the combination of the actions.
Under normal operating conditions, two or more load types will act on a structure
at any given time. In other words, the load types combine to produce more severe
conditions than if only single loads were to act. When this is considered, together
with the different stochastic characteristics of the various loads, it is not
reasonable to expect that all loads will exert their maximum lifetime values
simultaneously on the structure. The governing load effect due to a certain
combination of load types is found when one of the loads attains its lifetime
maximum value, and all of the other loads take on their arbitrary point-in-time
67

Design methods. Actions

values. Different methods can be used to account for the reduced probability of
heaving two or more loads appear simultaneously at their maximum lifetime
values. The effects are covered through the use of combination factor or of
different load factors for the same load type, depending on the particular
combination.
In theory, with the relatively large number of load types that may act on a
structure, the number of potential load combinations will be very large. However,
the only realistic design situations will be analysed in design process.

3.3. DESIGN STEPS OF WOOD STRUCTURES


3.3.1. Fundamental requirements
A wood structure shall be designed and constructed in such a way that:
with acceptable probability, it will remain fit for the purpose for
which it is required, being regard to its intended life and its cost;
with appropriate degrees of reality, it will sustain all actions and
influences likely to occur during execution and use and have
adequate durability in relation to maintenance cost.
A wood structure shall also be designed so that it will not be damaged by
events like explosions, impact, earthquake or consequences of human errors, to an
extent disproportionate to the original cause.
These requirements shall be met by the choice of suitable timber products,
by appropriate design and detailing and by specifying control procedures for
production, construction and use as relevant to the particular project.
For wood engineering structures design the Romanian design rules are given
by the standard STAS 856-71 and by code NP005-96. They are also based on
different prescriptions presented in other different units.

3.3.2. Design using allowable strength method


The Romanian standard STAS 856-71 is based on the allowable strength
method. The standard STAS 856-71 in tables 1 and 4 gives the values of
allowable strengths and displacements for different species of wood, and typical
loading. In chapter 2, Table 2.4 gives information regarding the allowable
strengths.
The standard allowable strength values are given for 15% moisture content.
Therefore, when the moisture content is more than 15% during lifetime of wood
68

Timber structures - 3

elements, or other parameters are different from those taken into consideration at
the test time, it is necessary to adjust them by other factors presented in standard
clauses.
The design process of a timber member and structure has the following
stages using ASDM (allowable strength design method):
(1) To determine the design values of the action effects (for examples
internal forces and moments). The procedure is based on the
characteristic values of the load (actions), the load arrangements and
load cases, and structure geometrical data:
- the characteristic values of the load are given by STAS 10101;
- a load arrangement identifies the position, magnitude and
direction of an action;
- a load case identifies compatible load arrangements, sets of
deformations and imperfections considered for a particular
verification. For each load case, design values for the effects of
actions shall be determined from combination rules. Wood
engineering design is based on three types of load combination.
They are:
I - sum of permanent actions + sum of long term variable
actions + one of short term variable actions;
II - sum of permanent actions + sum of long term variable
actions + two of short term variable actions;
III - sum of permanent actions + sum of long term variable
actions + sum of short term variable actions + accidental
actions.
These load combination rules must be analysed from the
probability point of view that the actions could apply together.
These rules are available for wood structures in general,
exception is met for wood roof beams in hypothesis I, where only
the maximum value obtained from the following load
combinations is valid:
I-1 - sum of permanent loads + snow loads;
I-2 - sum of permanent loads + wind loads + 1/2 snow loads;
69

Design methods. Actions

I-3 - sum of permanent loads + Q;


where Q = 8001000 [N] is a concentrated force given by the
weight of a person. (For more details one must refer to the
standard STAS 856-71).
(2) To establish the geometrical data describing the wood elements, which
are the cross-sectional dimensions, area, section modulus and the first
and the second moment of area. The timber element cross-section sizes
are generally represented by their nominal values.
(3) To evaluate the maximum effective stresses and displacement based on
the action effects and geometrical data.
(4) To select the material properties. They are represented by the
characteristic value, which, in general, corresponds to a fraction in the
assumed statistical distribution of the particular property of the specified
conditions of the material. The allowable value of the property of a
material is generally defined as its characteristic value divided by its
unique safety factor.
(5) To verify the equations (3.2). The allowable strength and displacement
values for wood structural design process are given by STAS 856-71. If
the conditions presented in eqs. (3.2) are not satisfied, it is necessary to
modify the geometrical characteristics of the timber elements and/or
structure.

3.3.3. Design using limit states method


The Romanian code NP005 96 and the European code EUROCODE 5
(EC 5) present the design procedure of timber structures and members based on
limit states method. According to LSDM (limit states design method) the design
process of a timber member and structure has the following stages:
(1) To determine the design values of the action effects:
- the characteristic values of the load are given by STAS 10101;
- for each load case, design values for the effects of actions shall be
determined from combination rules presented by equations (3.6)
& (3.7). The load and combination factors are given in standards.
(2) To establish the geometrical data describing the wood elements. The
procedure is the same for both methods.
70

Timber structures - 3

(3) To chose the wood strength class used in design process and to establish
the material properties, which are represented by the characteristic
strength and stiffness values. They are shortly presented in chapter 2,
Tables 2.5 & 2.6, and extensively in NP005-96, EC 5 or SR-EN 338.
(4) To determine the design resistance, which represents the load capacity
of timber element, using the equations (3.11; 3.12). The design
resistance, Rd, is given by:
R
(3.13)
R d k mod i a i mTi Ric a i mTi
i

where: - Rd = design resistance (load capacity);


- Ri = the characteristic strength of timber member subjected to
internal forces and moments as tension, compression,
bending, shear, etc. Using EC 5, the notation is replaced by
fk,i, which is also the characteristic strength (subscript i
means the type of loading);
c
- R i = the calculation strength value of timber member subjected
to internal forces and moments like tension, compression,
bending, shear, etc;
- ai = the geometrical characteristics (area, section modulus, etc);
- mTi = the coefficient due to the use of preservative substances;
its values range from 0.70 to 1.00 (table 4.1 of code
NP005-96).
- i = the partial safety factor for the specific property of the
material. Its values are given in table 2.7 in the Romanian
code NP005-96. A summary is presented here in Table 3.3.
Table 3.3
Type of load
Bending
Tension
Compression
Shear parallel to the grain
Shear perpendicular to the grain

Symbol
i
t
c
f//
f

values

1.10
1.20 1.40
1.25
1.10 1.25
1.10

The modification factor, kmod, which is taking into account the effect of
the duration of the actions and the moisture content, is given by the eq.
(3.10). The coefficient values for mui and mdi are given by the same
code in tables 2.5 and2.6.
(5) To verify the equations (3.3), (3.4) and (3.5). If they are not satisfied, it
is necessary to modify the geometrical characteristics or the strength
class of the timber elements.
71

Design methods. Actions

Try to
answer
these

Questions for you

1. Describe the basic principle of the allowable strength method.


2. Describe the basic principle of the limit states method.
3. Define the following terms: ultimate limit states and serviceability limit
states.
4. List the type of actions.
5. Describe the design process stages based on the allowable strength method.
6. Describe the design process stages based on the limit states method.
7. Make a comparison related to the design philosophy between the allowable
strength method and the limit states method.

72

Timber structures - 4

4.1. INTRODUCTION
The effects of the actions on a cross-section of an element are one or more
types of internal forces or moments. From the analysis of the loaded structure or
the loaded elements, the designer obtains the values of internal forces or moments,
which are available on a cross section or a node. The types of the internal forces
and moments are:
- tension force;
- compression force;
- shear force;
- bending moment;
- torsion moment.
The designer has to know the procedure to evaluate the maximum effective
(or actual) stresses and displacements. These values must be compared to the
permissible strengths or displacements given by the codes (standards) following
one of the method procedures presented in chapter 3 (allowable strengths design
method - ASDM or limit states design method - LSDM).
Beams are those members settled down in horizontal position. Usually they
are subjected to bending, but they could also be subjected to any type of the
internal forces presented above.
The design process of wood beams generally follows the same basic overall
procedure used in the design of other structural material beams. Although the
design principles are essentially the same, the material characteristics are different
(steel for example is ductile, homogeneous, and isotropic, concrete is brittle and
can be assumed homogeneous for most practical purposes). As for wood, the
properties of the material are different in the two main directions: parallel and
perpendicular to the grain.
73

Wood beam design

Beam sections are generally chosen on the basis of bending and then
checked for other possible failure modes. If not braced to prevent their
compression surface from moving sidewise, wood beams may buckle laterally due
to instability. Fortunately, most wood beams and joists are braced at close enough
intervals along their compression face that lateral buckling is prevented.
The following paragraphs deal in detail with general considerations and
present the checking formulas necessary for the design of wood beams. The rules
given by the Romanian standard STAS 856-71 and code NP005-96 are presented
for both methods: ASDM and LSDM. Design methods for glued laminated beams
and ply-webbed beams are described in chapter 7 and 8 respectively.

4.2. DESIGN OF RECTANGULAR WOOD BEAMS


The rectangular wood beams are those with a rectangular cross-section and
made of one piece of wood, fig. 4.1.
y

Figure 4.1 Cross-section of a rectangular


wood beam
b

Beam sections are specified by nominal dimensions, but cross-sectional


properties such as area and section modulus are always based on actual
dimensions.
Beam design consists of selecting a species, grade, and cross-sectional
dimensions to prevent any of these types of failure from occurring. Beam failure
could be a result of bending, shear, lateral buckling, bearing, or deflection under
service loads.

4.2.1. Member subjected to axial tension force


The wood member is loaded with an axial tension force (fig. 4.2). The
capacity of a tension member at its weakest point should be determined.
74

Timber structures - 4

Nt

Nt

Figure 4.2 Wood member subjected to axial tension force


Thus the maximum effective stress could be evaluated as:
Nt
An

max,ef

[N/mm2]

(4.1)

where: - Nt = axial tension force, [N];


- An = nominal value of cross-sectional area;
An Ab
A
[mm2]
- Ab = gross area of cross-section;
- A = the area of discontinuities (the sum of the discontinuity areas on a
200 mm length of the beam);
- max,ef = maximum effective tension stress.
The strength of the wood member is checked with:
ASDM:
max,ef

where: -

at

(4.2)

at

= allowable tension strength given in STAS 856-71, table 1.

LSDM:
Nt

Tr

Rdt

Rtc An mTt

k mod,t

Rt

An mTt

[N]

(4.3)

where: - Rdt = design tension resistance, [N];


- Rt = characteristic tension strength, [N/mm2], the values are given by
code NP005-96, table 2.3;
- kmod,t , mTt = factors for tension described in chapter 3;
- t = the partial safety factor for the property of the material for tension.

75

Wood beam design

4.2.2. Beam subjected to bending moment (flexural beam)


The flexural strength of a beam is generally the primary structural
consideration of the designer. The main design considerations for flexural beam
(fig. 4.3) are bending stress, shear stress, deflection, and bearing stress. This
concept can be stated simply by saying that any beam design should first be
approached by providing adequate flexural capacity, eq. (4.4). In typical
horizontal beam, vertical shear stress results from vertically placed loads. This
shear stress is not a failure mode in wood flexural members. However, vertical
shear is always accompanied by horizontal shear stress that is parallel to the grain
of the wood and must be considered in design. For the wood beam used as a
simple-span member, the horizontal shear is maximum at the supports and at near
the neutral axis, becoming zero at the top and bottom faces of the member. The
theoretical formula to determine horizontal shear stresses is given by eq. (4.5).
After this basic task has been satisfied, other design considerations need to
be taken into account and the art of wood beam design begins.

P
z

Figure 4.3 Flexural wood beam

4.2.2.1. Design for bending and shear


The design of wood beams in flexure requires the application of the elastic
theory of bending, expressed as:

max,ef

max,ef

M max
Wn
T max S
bI

[N/mm2]

(4.4)

[N/mm2]

(4.5)

where: - (M)max = maximum value of bending moment, [Nmm];


- (T)max = maximum value of shear force [N];
76

Timber structures - 4

- Wn = nominal value of section modulus about x-x axis, [mm3]:


I xx
h
y
Wn W xn
;
y
2
- S = first moment of area about x-x axis (major axis) [mm3];
S

bh 2
8

hh
24

- b = width of beam section, [mm];


- h = depth of beam section, [mm];
- I = second moment of area about x-x axis, [mm4];
I

I xx

bh 3
12

= maximum effective (actual) bending stress;


max,ef = maximum effective (actual) shear stress.
max,ef

The checking formulas for bending and shear are:


ASDM:
max,ef
max,ef

where: -

ai
af

(a)

(4.6)

(b)

= allowable bending strength, [N/mm2];


2
af = allowable shear strength, [N/mm ].
ai

LSDM:
M

max

Mr

Rdi

RicWn mTi

k mod,i

Ri mTiWn

[Nmm]

(4.7)

[N]

(4.8)

max

Lr

Rdf

R cf //

bI
mTf
S

k mod, f

R f // mTf bI
f //

where: - Rdi = design bending resistance, [Nmm];


- Rdf = design shear resistance, [N];
- Ri = characteristic bending strength, [N/mm2];
- Rf// = characteristic shear strength, [N/mm2];
- Mr = resisting moment;
- Lr = resisting shear force;
- kmod,i , mTi = factors for bending;
- kmod,f , mTf = factors for shear parallel to the grain;
- i = the partial safety factor for the material property for bending;
- f// = the partial safety factor for the material property for shear.
77

Wood beam design

4.2.2.2. Design for bearing


The ends of the beam may bear directly on a supporting member, which
could be either another wood member (fig. 4.4), a masonry wall, or a steel hanger,
girder, or bearing plate. A beam may be also stressed in bearing when a member
above transfers its load to the beam. Provision of adequate bearing area supports is
the design element requiring the least concern of the designer. The bearing
stresses in wood beams are developed due to compressive forces applied in a
direction perpendicular to the grain and occur in positions such as points of
support or applied concentrated loads. A bearing failure is not a collapse; it is
merely a crushing of fibres and generally not a serious failure.

b
bearing area

Figure 4.4 - Beam bearing

lp

The applied bearing stress is calculated with the following equation:


F max
[N/mm2]
(4.9)
max,ef
Abearing
where: - (F)max = bearing force, maximum reaction or concentrated load, [N];
- Abearing = bearing area = bearing length x width of the section, [mm2];
- max,ef = maximum effective bearing stress.
The verifications for bearing stresses are:
ASDM:
max,ef

where: -

as

(4.10)

as

= allowable bearing strength perpendicular to the grain.

LSDM:
F

max

Qr

Rdc

Rcc Ac mTc mr

k mod,c

Rc

Ac mTc mr

[N]

(4.11)

where: - Rdc = design resistance for compression perpendicular to the grain;


- Rc = characteristic compression strength perpendicular to the grain;
- kmod,c , mTc = factors for bearing described in chapter 3;
78

Timber structures - 4

= the partial safety factor for the material property for compression
perpendicular to the grain;
- mr = support coefficient. According to NP005-96 (clause 4.4.2),
this coefficient has values between 1.00 and 2.00 depending on the
ratios of wood element area and support area.
c

Sometimes bearing stresses are applied at an angle (neither parallel nor


perpendicular), fig.4.5. In this case, one needs to use Hankinsons formula, which
accounts for the strength of wood being greater for end-grain bearing than for
compression perpendicular to the grain.

Figure 4.5 Angle type of beam


bearing

4.2.2.3. Design for deflection


Besides having strength, a beam should also have adequate stiffness so that
it will not deflect or sag too much. Excessive deflections can lead to broken
windows, cracked plaster, ponding of water, or just bad appearance. Architectural
or structural damage is not the only reason for limiting deflection. Also, human
factor needs to be considered. People feel uncomfortable walking across a floor
that is saucer-shape, no matter how safe and strong the floor is.
The theory of beam action provides the basis for calculating the theoretical
deflection or vertical movement under load of a wood beam, in either a simplespan design or the more involved cantilever or continuous-span designs. This
theory assigns part of the deflection to the elongation of the wood fibres due to
bending, part to shear deformation and part to joints or support deformations. This
procedure in flexural beam is known as the verification of the rigidity condition.
ASDM:
The procedure requires that the final deflection under loading must be
limited by the allowable displacement and the verification equation is:

f max,ef

f adm

[mm]

(4.12)
79

Wood beam design

where: - fmax,ef = deflection under loading;


- fadm = permissible (allowable) deflection.
The equation (4.12) is checked, using the allowable deflection values
presented in table 4 of STAS 856-71. The loading deflections are evaluated
according to the end restraints of the beam and the load types (distributed or
concentrated loads).
LSDM:
The procedure follows the same formula presented in eq. (4.12), using the
allowable deflection values given in table 3.1 of code NP005-96. The evaluation
method of deflections under loading is modified such as it is presented in the
following equation:

f max,ef

max, final

f1

f2

f3

fc

[mm]

(4.13)

where:
- f1 = deflection of the beam due to the permanent loads immediately after
loading (state 1);
- f2 = deflection of the beam due to variable loads plus any time dependent
deflection due to permanent loads (state 2);
- f3 = deflection due to joint deformations (state 2);
- fc = precamber of the beam in the unloaded state (state 0).
Beam deflections f1 and f2 are evaluated with eq. (4.14):

fi

f i ,init 1 k def

i 1,2

(4.14)

where: - fi,init is evaluated with theoretical formulas according to the end restraints
and loading types;
- kdef = deformation factors, having values between 0.00 and 1.00.
The values of deformation factors are presented in table 3.2 of
NP005-96.
Values for f3 are given in NP005-96, table 3.3 and they vary according to
the joint types.

4.2.3. Beam subjected to axial tension and bending moment


In members that are subjected to transverse (lateral) loading as well as the
axial tension, fig. 4.6, the position of maximum stress occurs at the point of
maximum bending.
80

Timber structures - 4

P
Nt

Nt

Figure 4.6 Tension member subjected to lateral loading


The actual stress is determined with:

max,ef

Nt
An

M max
Wn

ef ,t

max,ef ,i

[N/mm2]

(4.15)

where: - Nt = axial tension force, [N];


- (M)max = maximum bending moment, [Nmm];
- An = nominal value of cross-sectional area;
- Wn = nominal value of section modulus about x-x axis;
- ef,t = effective tension stress, [N/mm2];
- max,ef = maximum effective bending stress, [N/mm2].
The sign
is due to the fact that bending stress could be tensile or
compression stress, and the maximum stress value is obtained when both types of
stresses (axial and bending) have the same sign.
The following condition needs to be satisfied:
ASDM:
at
ef ,t

max,ef ,i

at

(4.16)

ai

where: -

= allowable bending strength;


at = allowable tension strength.
ai

LSDM:
Nt
Tr

M max
Mr

(4.17)

where Mr and Tr are calculated according to eqs. (4.7) and (4.3).


81

Wood beam design

4.2.4. Beam subjected to bi-axial bending


P

Figure 4.7 Beam subjected


to bi-axial bending

Mx
My
b

The most common use of a beam is to resist loads by bending about its
major principal axis. However, the introduction of forces, which are not in the
plane of bending on the beam, results in bi-axial bending. (i.e. bending about both
the major and minor principal axes). In fig.4.7, the case when the loading direction
is different than the principal axis directions is presented. The load direction is
also passing through the centroid and it produces bi-axial bending.
In these cases, bending stresses induced in the member become additive.
Therefore, the maximum effective bending stress is a combination of bending
moment about x-x and y-y axes and equates to:
Mx
max,ef

max

W xn

My

max

W yn

[N/mm2]

(4.18)

where: - (Mx)max = bending moment about x-x axis (major axis), [Nmm];
- (My)max = bending moment about y-y axis (minor axis), [Nmm];
- Wxn = nominal value of section modulus about x-x axis, [mm3];
I xx
h
W xn
; y
y
2
- Wyn = nominal value of section modulus about y-y axis, [mm3];
I yy
b
; x
W yn
x
2
- Ixx, Iyy =second moments of area about x-x and y-y axes respectively.
I xx

bh 3
;
12

I yy

hb 3
12

Depending on the method used, one of the following conditions must be


verified:
82

Timber structures - 4

ASDM:
max,ef

(4.19)

ai

where ai is the allowable bending strength. For solid sawn members, the
allowable bending strengths in both the x-x and y-y axes are assumed to be the
same.
LSDM:
Mx

My

max

max

1
Mx r
My r
where (Mx)r and (My)r are evaluated with eq. (4.7).

(4.20)

The rigidity condition expressed as a limited deflection is verified with the


equations:
ASDM:
x
f max,
ef

f max,ef

where f xmax,ef and f


respectively.

y
max,ef

y
f max,
ef

(4.21)

f adm

are the deflections about the x-x and y-y axes

LSDM:
x
f max,
final

y
f max,
final

(4.22)
where the final deflections, f xmax,ef and f ymax,ef are evaluated with the procedure
presented in paragraphs 4.2.2, eqs. (4.13) and (4.14).
f max, ef

max, final

f adm

This type of bending stress is generally met at roof purlin beams, which are
installed normal to the slope of the roof. The successful design of this type of
beam is largely a question of trial and error. Wherever possible a purlin beam
should be placed in the vertical plane and if necessary the supported construction
should be notch seated over. This simplifies the design very much and gives an
economical section. However, where purlins are placed normal to the roof slope,
careful attention must be given to the end support where possible sliding and
rotation can occur.

4.3. DESIGN OF FLITCHED BEAM


4.3.1. Flexural beam
Very often in wood design there arises the situation where not only large
span and heavy loads predominate but also the available depth of the section is
restricted in some way. When this does happen, the more conventional choice of
83

Wood beam design

sections invariably fails to produce a solution. Consequently, it is not unusual for


the designer to investigate the introduction of a flitched beam.
Flitched beam is a composite beam formed by sandwiching a steel plate
between two wood beams and bolting together at intervals (fig.4.8).
wood members

h1

x
steel plate

b
y

Figure 4.8 Flitched beam cross-section


The design procedure is presented for allowable strength methods (ASDM).
If we note the ratios of elasticity modulus of wood and steel with k=Es/Ew
the verification equation for moment capacity of the section is:

M
where: -

w
ai

Wnw

max

w
ai

Wns

[Nmm]

(4.23)

k = modular ratio;
w
2
ai = allowable bending strength for wood, [N/mm ];
Wnw = nominal value of wood section modulus (Wnw = bh2/3);
Wns = nominal value of section modulus for steel (Wns = th12/6);
Ew = elasticity modulus of wood material, [N/mm2];
Es = elasticity modulus of steel, [N/mm2].

The verification equation for shear capacity is:


T

max

2
3

w
af

Anw

w
af

Ans

[N]

where: - Anw = 2bh = nominal value of wood area, [mm2] ;


- Ans = th1 = nominal value of steel area, [mm2];
- afw= allowable shear strength for wood, [N/mm2].
84

(4.24)

Timber structures - 4

The deflection is checked with eq. (4.12) and using the stiffness:
EI

E wI w

EsI s

(4.25)

where: - E = composite elasticity modulus, [N/mm2];


- Ew = elasticity modulus of wood material, [N/mm2];
- Es = elasticity modulus of steel, [N/mm2];
Iw

bh3
12

bh3
= second moment of area for wood members
6

about x-x axis, [mm4];

th13
= second moment of area for steel plate about x-x axis, [mm4].
12

4.4 DESIGN OF NOTHCHED BEAMS


In figure 4.9 beams with various types of notches are shown. A notch may
very significantly reduce the load bearing capacity of the beam and should be
avoided in design. Though not to be desired, a notch may be needed in order to
bring floors to desired levels, to give clearance or to enable fit between structural
members.

(a)

(b)

(c)

Figure 4.9 Notched beams


Fracture may develop from a notch by the broken lines in fig. 4.9. The
fracture is often of a very sudden and brittle nature, taking place without being
preceded by any large deformation or after visible warning. Depending on the
geometry of the beam, the rapid crack propagation along the beam may or may
not lead to a complete collapse of the beam.
The initiation of crack growth is due to perpendicular to grain tensile stress
or shear stress or a combination of the two. At the tip of a notch these stresses may
become very high. According to linear elastic stress analysis the stress at the tip of
a sharp notch even approaches infinity, fig. 4.10, [27].
Due to the limited strength of the material the stress at the tip of the notch
does, in reality, not approach infinity. Instead, due to local damage of the material,
85

Wood beam design

the stress distribution at the instant crack starts to propagate as indicated by the
broken curve in fig. 4.10.
t,90

ft,90

i(1- )h

Vd
h

Figure 4.10 Stress at the tip of a notch according to linear elastic theory and
as estimated in practice, respectively
The risk of crack propagation from a notch is taken into account in EC5
through a formal reduction by a factor kv of the design shear strength, fv,d, of the
net cross section b h:
d

3Vd
2b h

(4.26)

k v f v ,d

where the reduction factor kv (<1.0) is:


k n 1 1.1

i 3.5

kv
h

0.8

(4.27)
1
2

where i is notation according to fig. 4.10.


For solid timber kn must be equal to 5.0 and for glulam it is set equal to 6.5. Note
that the beam depth, h, must be in [mm]. To avoid the risk of the development of
another mode of failure, shear failure in the net cross-section, a value of kv greater
than 1.0 may not be used in eq. (4.26). If the notch is located on the compression
side of the beam (fig. 4.7.c) kv may be set equal to 1.0.

4.5 DESIGN EXAMPLES


4.5.1 Design of a main beam
A main beam of 3.45 m length spans over an opening 3.3 m wide (fig. 4.11)
and supports a flooring system which exerts a long-duration loading of 3.5 kN/m,
86

Timber structures - 4

including its own self-weight, over its span. The beam is supported by 50 mm
wide walls on either side. Carry out design checks that a 100 mm x 200 mm deep
sawn section fir under service class 1 is suitable.

200

100
3300 mm
50 50

50 50

3250 mm

Figure 4.11 Main beam - example

Design method: ASDM STAS 856-71


Wood
Geometrical properties:
Cross-section dimension 100 x 200 mm;
Cross-sectional area, A = bh = 20000 mm2;
Second moment of area, Ixx = bh3/12 = 6.67 x 107 mm4;
Span (clear distance), L = 3250 mm;
Bearing width, bw = 50 mm;
Effective span, Le = L + bw = 3300 mm.
Mechanical properties:
Species fir tree;
Moisture content 15%
Allowable bending strength ai = 10 N/mm2 ;
Allowable shear strength parallel to the grain
Allowable bearing strength as = 2.5 N/mm2;
Elasticity modulus E = 10000 N/mm2;
Allowable deflection fadm = Le/250.

af//

= 2 N/mm2 ;

Loading:
Applied uniformly distributed load, p = 3.5 kN/m
Load duration class long term;
Service class 1.
87

Wood beam design

Static scheme:
Simply supported beam with uniformly distributed load, span = Le.
Bending stress
Applied maximum bending moment:
pL2e 3.5 3300 2
[Nmm]
M max
4764375
8
8
Section modulus (it is supposed there are no defects and reductions on the crosssection):
I xx I xx bh 2
666666.67 [mm3]
Wn
h
6
y
2
Maximum effective bending stress, eq. (4.4):

7.15
[N/mm2]
The equation (4.6.a) is checked.
max,ef

Shear stress
Maximum effective shear stress, eq. (4.5):
[N/mm2]
0.433
max,ef
The equation (4.6.b) is checked.
Bearing stress
Bearing stress is:
1.155
[N/mm2]
max,ef
Abearing = 50 x 100 =5000 [mm2];
Bearing condition is verified.
Deflection
The deflection is:
4

5 pLe
f ef
8.1
384 EI xx
Deflection is satisfactory.

f adm

13.2

[mm]

4.5.2 Design of floor joists


A wood floor spanning 3.5 m centre to centre is to be designed using wood
joists at 400 mm centres, fig. 4.12. The floor is subjected to a domestic imposed
load of 1.5 kN/m2 and carries a dead loading, including self-weight of 0.40 kN/m2.
88

Timber structures - 4

Carry out design checks to show that a series of 50 mm x 100 mm deep sawn
section oak under service class 1 is suitable for the verification of eq. (4.7).
A

3500 mm

Main beam

Joist
Tongued & grooved boarding
Joists
x

400 mm

400 mm

400 mm

Plasterboard

Figure 4.12 Floor joist - example


Design method: LSDM NP005-96
Wood
Geometrical properties:
Cross-section dimension 50 x 100 mm;
Cross-sectional area, A = bh = 5000 mm2;
Second moment of area, Ixx = bh3/12 = 4.17 x 106 mm4;
Span (clear distance), L = 3500 mm;
Bearing width, bw = 100 mm;
Effective span, Le = L + bw = 3600 mm.
Mechanical properties:
Species fir tree;
Moisture content 15%
Characteristic bending strength, Ri = 24.0 N/mm2;
Characteristic shear strength, Rf// = 3.0
Modification factor for bending:
- for dead load, k1mod,i = mui x mdi = 1.00 x 0.55 = 0.55
- for domestic imposed load, k2mod,i = mui x mdi = 1.00 x 0.65 = 0.65
Bending factor due to the preservative substances, mTi = 1.00 (no treated wood);
Partial safety factor for the material property for bending, i = 1.10;
89

Wood beam design

Loading:
Applied uniformly distributed load: p = 0.4 x (1.50 + 0.40) = (0.6 + 0.16) kN/m
- domestic imposed load of 1.5 kN/m2;
- dead loading, including self-weight of 0.40 kN/m2
Load duration class long term;
Service class 1.
Static scheme:
Simply supported beam with uniformly distributed load, span = Le.
Bending stress
Applied maximum bending moment:
M

pL2e
8

max

0.6 3600 2

0.16 3600 2
8

972000 259200 1231200

[Nmm]
Section modulus (it is supposed there are no defects and reductions on the crosssection):
Wn

I xx
y

bh 2
6

I xx
h
2

83333.33

[mm3]

The equation (4.7) is checked:


M r1
M r2

Mr

1
k mod,
i

2
k mod,
i

1
r

Ri mTiWn

0.55

Ri mTiWn

Ri mTiWn

0.65

Ri mTiWn

i
2
r

999999.95

[Nmm]

1181818.12 [Nmm]

2181818.07 [Nmm]

Equation (4.7) is satisfied.

4.5.3 Axial load capacity of a tensile member with lateral load


A trussed rafter tie (fig. 4.13) of 40 mm x 100 mm section is 2.7 m long
and is subjected to a lateral concentrated load of 0.8 kN at mid-length. Determine
the maximum axial tensile load that the rafter tie can carry. Assume a wood
member made of spruce under service class 2 conditions.
90

Timber structures - 4

0.8 kN
T

0.4 kN

L = 2.7 m

0.4 kN

Figure 4.13 Trussed rafter tie - example

Design method: LSDM NP005-96


Wood
Geometrical properties:
Cross-section dimension 40 x 100 mm;
Cross-sectional area, A = bh = 4000 mm2;
Second moment of area, Ixx = bh3/12 = 3.33 x 106 mm4;
Effective span, Le = 2700 mm.
Mechanical properties:
Species fir tree;
Moisture content 15%
Characteristic bending strength, Ri = 24.0 N/mm2;
Characteristic tension strength, Rt = 14.0
Modification factor for bending:
- for dead load, kmod,i = mui x mdi = 1.00 x 0.55 = 0.55
Bending factor due to the preservative substances, mTi = 1.00 (no treated wood);
Modification factor for tension:
- for dead load, kmod,t = mut x mdt = 1.00 x 0.90 = 0.90
Tension factor due to the preservative substances, mTt = 1.00 (no treated wood);
Partial safety factor for the material property for bending, i = 1.10;
Partial safety factor for the material property for tension, t = 1.20.
Loading:
Lateral concentrated load P = 0.8 kN at mid-length;
Load duration class long term;
Service class 1.
Static scheme:
Simply supported beam with a concentrated load, span = Le.

91

Wood beam design

Tension capacity
Tr
Nt
M max
Nt
Mr
1
Mr
Tr
Mr
Applied maximum bending moment is:

pLe
540000 [Nmm]
4
and the resistant force and moment are:
Rm W
Rm W
M r k mod,i i Ti n 0.55 i Ti n
M

max

max

Tr

k mod,t

800000.04

[Nmm]

Rt mTt An

0.90

Rt mTt An

43200

[N]

Therefore Nt is evaluated as:


Nt

Tr
Mr
Mr

max

14040

[N]

The maximum value of tension force could be chosen 14 kN.

Try to
answer
these

Questions for you

1. Compare the results of the above examples, which occur in the design
processes using ASDM and LSDM.
2. Give and solve an example for bi-axial bending.
3. Develop the conditions of a notched beam and design the beam subjected to
a uniformly distributed load in all the cases presented in fig.4.9. Make a
comparison for both design methods.
92

Timber structures - 5

5.1. INTRODUCTION
Columns are generally thought of as the vertical supporting members of
buildings. However, there are other structural members that act as columns: the
piers of a bridge or the compression chord of a truss. Generally columns are
compression members, but they can also have combined compression and bending
or even tensile axial force under loading that cause uplift. For purposes of design,
we define a column as a structural member whose primary loads are axial
compression. Compression members include posts or columns, vertical wall studs,
and struts in trusses and girders.
This chapter minutely deals with the general considerations necessary for
the design of compression members. For this part, we will assume that a prop
column is defined as a vertical piece of timber pinned or fixed at each end,
carrying a load applied at the upper end, as they are indicated in fig. 5.1.
N

N
Nu

Nu

(a)

(b)

Figure 5.1 Two-hinged column (a) and pinned-fixed column (b)

93

Wood column design

When a slender column is loaded axially, there exists a tendency for it to


deflect sideways. This type of instability is called flexural buckling. The strength
of slender members depends not only on the strength of the material but also on
the stiffness, in the case of timber columns mainly on the bending stiffness.
Therefore, apart from the compression and bending strength, the modulus of
elasticity is an important material property influencing the load-bearing capacity
of slender columns. There are two principal ways to design a compression
member: the first involves a second order analysis whereby the equilibrium of
moments and forces is calculated by considering the deformed shape of the
respective member of structure. The second approach uses buckling curves to
account for the decrease in strength of a real column compared to a compression
member, which is infinitely stiff in bending. Here, the stability design is carried
out as a compression design with modified compression strength. The decrease in
load-bearing capacity depends on the slenderness ratio of the member in question.
Wood columns are available in a variety of shapes and types. Most often
they are either rectangular, square, or round in cross section. They may be made
of a single solid piece of wood or formed from several pieces of wood.
The following paragraphs deal with general considerations and the rules
given by Romanian standard STAS 856-71 and code NP005-96 are presented for
both methods: ASDM (allowable strength design method) and LSDM (limit states
design method). Design procedures for glulam columns and columns of simple
composites will be described in the next chapters.

5.2. END RESTRAINTS AND EFFECTIVE LENGTH


We know that the deflected shapes of buckled columns will differ
depending on the end support conditions. In most cases the effective length will
differ from the actual unbraced length. Figure 5.2 shows the theoretical design
values of effective lengths.
The correct interpretation of the end restraints is an important feature of the
design because, as will be shown, it governs the chosen effective length, the
slenderness ratio and, consequently, the ultimate load, which may be carried by
the column. Determination of end restraints is a matter of individual interpretation
primarily born of much variation in experience. The assumptions of one designer
may not necessarily satisfy another and so, having chosen, the designer should be
prepared to defend his choice with sound reasoning. In the event of doubt, then
always choose pinned at both ends and take the full height of the column as its
effective length.
Generally end restraints fall into categories:
positional - held in line but free to rotate, i.e. pinned;
positional and directional - held in line and against rotation, i.e. fixed.
94

Timber structures - 5

lf

lf

lef

lf

l
2

lf

l
2

2l

Figure 5.2 Buckling lengths for idealised support


condition

The effective or buckling length of a compression member is defined as the


length of a hypothetical two-hinged column with the same elastic critical buckling
load as the member in question. The effective length can be visualised as the
distance between two consecutive points of contra-flexure of the actual
compression member. Figure.5.2 shows the theoretical effective (bucking) lengths
of a column, lf, with no lateral restraint in length. The last two values are different
in the case of wood members due to the shrinkage and bearing failure
perpendicular to the grain direction. These values are 0.80 and 0.65 instead of 0.71
and 0.50 respectively.
STAS 856-71 presents in tables 7 and 8 the effective lengths for all types of
columns, which can be met in wood engineering design. Code NP005-96 presents
in table 3.4 the effective lengths according to the types of column supports.

5.3. SLENDERNESS RATIO AND BUCKLING


COEFFICIENT
Columns of all materials are classified by their slenderness ratio.
Slenderness ratio, , is the ratio of the effective length, lf, to the radius of
gyration, i. The radius of gyration, i, can be calculated about x-x axis or y-y axis
and therefore the slenderness ratio may differ for the two principal axes of the
cross section. If so, the larger of the two slenderness ratios is the critical one.
Thus:
I xx
(5.1)
ix
Ac
and
95

Wood column design

I yy

iy

(5.2)

Ac

Therefore, slenderness ratios are:

l fx
x

ix
l fy

iy

(5.3)
(5.4)

and the critical slenderness ratio will be the larger of the two slenderness ratios:

max

(5.5)

where: - Ixx = second moment of area about x-x axis;


- Iyy = second moment of area about y-y axis;
- lf = effective length of the column;
- ix = radius of gyration about x-x axis;
- iy = radius of gyration about y-y axis;
- x= slenderness ratio about x-x axis;
- y= slenderness ratio about y-y axis.
- Ac = calculation area of the cross section, which has been evaluated
using the following rules according to the details presented in fig. 5.3:
-

Ac Ab , when the reduction area is less than 25% of the gross


area, Ab, and the reductions are on the faces perpendicular on
buckling direction, (fig. 5.3.a, b);
4 An
Ab , when the reduction area is greater than 25% of
3
the gross area, and the reductions are on the faces perpendicular
on buckling direction, (fig. 5.3.b). An is the nominal value of
cross-sectional area and An Ab
A where A is the area of
discontinuities;
Ac

Ac An , when the reductions are symmetric about buckling


directions, (fig. 5.3.c);

- in the case of an unsymmetrical reduction, fig. 5.3.d, the members


are designed subjected to eccentrical load.
96

Timber structures - 5

(a)
(b)
(c)
(d)
Figure 5.3 Types of common reductions in compression columns
If the unbraced length of a column is the same in both directions, the
column will buckle about the y-y axis (fig. 5.4). In other words its middle will
move in x-x direction. Because of the tendency to buckle about y-y axis, this axis
will be called the weak axis, and the x-x axis will be called the strong axis.
y

Figure 5.4 Column cross-section axes:


buckling weak and strong axes
y

If the unbraced length of the column is not the same about both the strong
and weak axes when considering buckling about these two axes of the section,
also the larger of the two slenderness ratios is used for design calculations, figure
5.5. In cases where buckling is effectively prevented about one axis, the
slenderness ratio to be taken in the design is determined by taking the unbraced
length for the other axis. Such a case can exist in studs in a bearing wall.
97

Wood column design

Figure 5.5 Slenderness ratios in


simple solid column
- slenderness ratio about axis y-y:
l
12 1
y
b

l1

l2

- slenderness ratio about axis x-x:


l
12 2
x
h

l1

Buckling coefficient,
(eq. (5.6)), is a modification factor for failure
strength giving the critical stress for a perfect elastic column:
crit

(5.6)

where

is the failure strength.

The critical compression stress is determined using the formula:


2
crit

E
A
I

l
i

E
2

(5.7)

where the Eulers buckling force is:

N crit

98

EI
l 2f

(5.8)

Timber structures - 5

As the slenderness ratio decreases, Eulers buckling stress increases,


tending to infinity as the slenderness ratio vanishes. This is clearly nonsense, since
the upper limit of the buckling stress is the yield strength of the material. Eulers
formula applies only to elastic buckling where the stress remains below the yield
point.
In practice, the buckling load is always lower than the theoretical Euler
value, due to imperfections such as lack of straightness of the member. The
deviation from Eulers load is largest in the region of slenderness ratios where
Eulers buckling load approaches the yield strength. The transition slenderness
ratio for wood is approximately 75. Columns, which possess a substantially lower
slenderness ratio than the transitional value, are termed stub columns and the
buckling coefficient is evaluated as:
2

1 0.8

when

100

(5.9)

75

It turns out that, in practice, the majority of axially compressed wood


members fall in the zone of slenderness ratios close to 75 150 and, in this
range, buckling conducives to progressive collapse. For slenderness ratio values
greater than 75, the following formula is used:

3100

when

75

(5.10)

Table 4.3 of the code NP005-96 gives the values of the buckling
coefficient according to the slenderness ratios and a summary of them is
presented here in Table 5.1 (the values of the first row and column give the
values of slenderness ratios)
Table 5.1
9
0.993

0
1.000

1
1.000

2
1.000

3
0.999

4
0.999

5
0.998

6
0.997

7
0.996

8
0.995

70
80

0.608
0.484

0.597
0.472

0.585
0.461

0.574
0.450

0.562
0.439

0.550
0.429

0.537
0.419

0.523
0.409

0.509
0.400

0.496
0.391

190
200

0.086
0.077

0.085
-

0.084
-

0.083
-

0.082
-

0.081
-

0.081
-

0.080
-

0.079
-

0.078
-

The buckling coefficients could also be evaluated using the diagram in


figure 5.6 according to slenderness ratio values.
99

Wood column design

1.00
0.90

0.80

1 0.8

0.70

100

0.60
0.50

3100

0.40

0.30
0.20
0.10

=75

0
20

40

60

80

100

120

140

160

180

200

Figure 5.6 Theoretical curve for buckling coefficient


versus slenderness ratio

5.4. DESIGN OF RECTANGULAR WOOD COLUMNS


Wood sections are commonly used in construction as axially loaded
members or members in combined axial force and bending, figures 5.7 and 5.8.
y

ex y

Load Nc

ey
x

ey

(b)

Load=Nc
M=Ncey

(c) Load=Nc
M=Ncex

Figure 5.7 Axial concentric and eccentric loads

100

(a) Load =Nc


M=0

ex y
(d) Load=Nc
Mx=Ncey
My=Ncex

Timber structures - 5

5.4.1. Column subjected to axial compression force


The element is loaded with an axial compression force, fig. 5.7.a. The
effective values must be compared with permissible strengths given by the codes
and standards following one of the methods procedures presented (allowable
strengths design method - ASDM or limit states design method - LSDM).
The maximum effective stress could be evaluated as:
Nc
[N/mm2]
(5.11)
max,ef
Ac
where: - Nc = axial compression force, [N];
- Ac = calculation value of cross-sectional area, [mm2]
(presented in subchapter 5.3);
- max,ef = maximum effective compression stress.
The strength and the stability conditions of wood member are checked
with:
ASDM:
max,ef

(5.12)

ac

Nc
(5.13)
ac
Ac
where: - ac = allowable compression strength given in STAS 856-71, table 1;
- = buckling coefficient evaluated with eqs. (5.9) or (5.10).

LSDM:
Nc

Cr

Rdc

Rcc //

Ac mTc

k mod,c

Rc

Ac mTc

[N]

(5.14)

where: - Rdc = design compression resistance;


- Rc = characteristic compression strength, the values are given by
code NP005-96, table 2.3;
- kmod,c , mTc= factors for compression described in chapter 3;
- c = the partial safety factor for the material property for compression.

5.4.2. Column subjected to axial compression force and moment


This loading condition can occur when the load is an eccentrically applied
one, where it acts at a given distance from the centre point of the section (see fig.
5.7.b, c, d) and thus it will induce moment about x-x and/or y-y axes. Another
case is met when the axial compression load is combined with bending moment
produced by side load, fig. 5.8. In the first case the fraction of stress which is
101

Wood column design

produced by the moment is constant along the length of the column, while in the
second it is variable along the length of column.
Nc

P
P
P
P
P

Figure 5.8 Column subjected to


compression plus bending
due to transverse loads

P
P
P
P

The formulas presented below are available for both cases: columns
subjected to eccentrical compression load and columns subjected to axial
compression load combined with transverse loads.
Members which are restrained at both ends in position but not direction,
which covers most real situations, should be so proportioned that:
ASDM:
max, ef

Nc
An

Nc
y Ac

M0
Wn

ac

ac

[N/mm2]

(5.15)

[N/mm2]

(5.16)

A [mm2]
where: - An = nominal value of cross-sectional area: An Ab
- Ab = gross area of cross-section, [mm2];
- A = the area of discontinuities, [mm2];
- Wn = nominal value of section modulus, [mm3];
- M 0 M max,ef N c f ;
- Mmax,ef = maximum effective bending moment, [Nmm];
- f = deflection due to bending moment, [mm];
- = coefficient taking into consideration the bending moment influence .
This coefficient has values between 0.00 and 1.00 and could be
evaluated with:
2

Nc
3100 Ab ac

- Ac = calculation value of cross-sectional area, [mm2], (subchapter 5.3);


102

Timber structures - 5

= buckling coefficient about y-y axis using eqs. (5.9) or (5.10).

LSDM:
f
N c M ef
1
Cr M r
where Mr and Cr are calculated according to eqs. (5.18) and (5.19):
Mr

Rdi

RicWn mTi

k mod,i

Ri mTiWn

(5.17)

[Nmm]

(5.18)

[N]

(5.19)

Cr

Rdc

Rcc //

Ac mTc

k mod,c

Rc

Ac mTc

The maximum value of the final bending moment, M eff , is given by eq.
(5.20):
M eff

M ef ,max

1
Nc
1
CE

[Nmm]

(5.20)

and
I
(5.21)
l 2f
where: - muE, mTE = working condition factors given in code NP 005-96,
tables 2.5 and 4.1;
- E0.05 0.67 E // and E// is the characteristic elasticity modulus parallel to
grain direction;
- I = second moment of area about axis at right angle as against the load
direction, [mm4];
- lf = effective length, [mm];
- Rdc = design compression resistance, [N];
- Rc = characteristic compression strength, [N/mm2], the values are given
by code NP005-96, table 2.3;
- kmod,c , mTc= factors for compression described in chapter 3;
- c = the partial safety factor for the property of the material for
compression;
- Rdi = design bending resistance, [Nmm];
- Ri = characteristic bending strength, [N/mm2] the values are given by
code NP005-96, table 2.3;
- kmod,i , mTi= factors for bending described in chapter 3;
- i = the partial safety factor for the property of the material for bending.
CE

E 0.05 mu ,E mTE

103

Wood column design

5.5. DESIGN OF COMPOUND COLUMNS (SPACED


COLUMNS)
Compound columns and their cross-sections are presented in fig.5.9.
Nc

Nc

Nc

Nc

l1

Nc
y
x

Nc Ap
y

Ap
1

Nc
y

Nc Ap
y

Ap

y
hp

(a)

hp

hp
hs

As

hs

(b)

As

hs

(c)

Figure 5.9 Compound columns

104

y
As
hp hp

hp

hp
hs

hs

hs

(d)

Timber structures - 5

Compound (spaced) columns are made of wood and steel members,


connected by nails, bolts or screws. They could be made of wood lumbers only, as
it is shown in fig.5.a. The steel plates, fig. 5.9.b,c,d, are used to increase the
stiffness of the compound column to prevent buckling about stack direction.
Therefore, they have a different buckling behaviour under axial compression load
combined or not with bending related to the main axes of the cross-section. The
slenderness ratio is evaluated taking into consideration the contribution of steel
plates, which are considered as secondary parts of the column. Wood lumbers are
the principal parts of the column structures.

5.5.1. Compound column subjected to axial compression force


The individual members in a compound column are considered to act
together to carry the total column load. The total load capacity, which is
determined by using the spaced column formulas, should be checked against the
sum of the load capacities of the individual members taken as simple solid
columns. Applying a modification procedure in determining the slenderness ratios
to those established for simple solid columns develops the design formulas for
compound (spaced) columns.
The same equations (5.12), (5.13) and (5.14) are used for checking the
strength and stability conditions (ASDM) or design compression resistance
(LSDM) considering that the nominal area is the sum of wood member crosssection areas, An
A p . The verifications are done about x-x and y-y axes,
because the column has different buckling coefficient values. The slenderness
ratios of compound columns, according to the type of cross-section are named
transformed slenderness ratios and are evaluated using the following
procedures:
Compound column fig. 5.9.a
l fx
l
tr
12 fx
x
ix
b
l fy
l fy
tr
12
y
y
iy
h

(5.22)
(5.23)

Compound column fig. 5.9.b,c


tr
x

l fx
ix

l fx
I px

0.5 I sx

(5.24)

Ap
105

Wood column design


tr
y

l fy
y

l fy

iy

I py

(5.25)

I sy
Ap

Compound column fig. 5.9.d


l fx
l fx
tr
12
x
ix
b

(5.26)

m 2
(5.27)
1
2
In equations (5.22), (5.23), (5.24), (5.25), (5.26) and (5.27) the new terms
tr
y

are:
-

coefficient taking into account the sizes of wood pieces, the number
of gaps, the number of shear sections and the mechanical fastener
types.
- Ipx = second moment of area about x-x axis for principal (wood) members;
- Ipy = second moment of area about y-y axis for principal (wood) members;
- Isx= second moment of area about x-x axis for secondary (steel) members;
- Isy = second moment of area about x-x axis for secondary (steel) members;
- m = number of wood members;
l
- 1 1 where l1 is distance between two consecutive spacing plates and i1
i1
is the radius of gyration for a single wood piece about 1-1 axis.
The coefficient is calculated according to specifications and more details
are given in standard STAS 856-71, clauses 7.3.2.1.b, 7.3.2.2.a, and 7.3.2.2.b
(ASDM) or code NP005-96, paragraphs 5.3 and 5.4 (LSDM).

5.5.2. Compound column subjected to axial compression force and


moment
The following formulas may be used for compound columns subjected to
both axial compression and moment (from eccentricity or side loads):
ASDM:
max,ef

or/and

Nc
An

My
k w y W yn

ac

[N/mm2]

Nc
Mx
[N/mm2]
ac
An
x W xn
where: - An = nominal value of cross-sectional area, [mm2];
max,ef

106

(5.28)
(5.29)

Timber structures - 5

- Wyn = nominal value of section modulus about y-y axis parallel to gaps;
- Wxn = nominal value of section modulus about x-x axis perpendicular to
gaps direction, [mm3];
- My = bending moment about y-y axis, [Nmm];
- Mx = bending moment about x-x axis, [Nmm];
- x ; y = coefficients taking into consideration the bending moment
influence . These coefficients are evaluated with:
x

tr 2
x

Nc
and
3100 Ab ac

tr 2
y

Nc
3100 Ab ac

- Ac = calculation value of cross-sectional area, [mm2], (subchapter 5.3);


- Ab = gross area of cross-section, [mm2];
- kw = reduction factor due to the joint deformations. The value 0.90 is
taken for one gap and 0.80 is taken for two or more gaps.
It is also necessarily to check a single element to buckling with the
relationship:
My
Nc
(5.30)
1 ac
Ab
1Wb
where: -

= buckling coefficient about 1-1 axis, evaluated for a wood member


with effective length l1 and radius of gyration i1;
- Wb = section modulus of gross area.

LSDM:
Nc
Cr
or/and
Nc
Cr

M eff , y
M rc, y

M eff ,x
M rc,x

(5.31)

(5.32)

where: - Nc = axial compression force, [N];


- Cr = design compression resistance, [N];
- M r , cy = design bending resistance of the column about y-y axis, [Nmm];
- M r , xc = design bending resistance of the column about x-x axis,[Nmm];
- M eff , y = the maximum value of the final bending moment about y-y axis,
- M eff , x = the maximum value of the final bending moment about x-x axis,
The evaluations of M eff , y and M eff , x are done according to eq. (5.20).
The design compression and bending resistances are:
107

Wood column design

Cr

Rdc

Rcc//

Ac mTc

k mod,c

Rc

Ac mTc

[N]

(5.33)

[Nmm]

(5.34)

[Nmm]

(5.35)

M rc,y

Rdic

k w RicW yn mTc

k w k mod,i

Ri mTiW yn
i

M rc,x

Rdic

RicWxn mTc

k mod,i

Ri mTiWxn
i

The verification formula of a single element to buckling is:


f
N c M ef
1
Cr M rc

(5.36)

For more specific details, the standard STAS 856-71 and code NP005-96
must be observed.

5.6. ROUND AND TAPERED COLUMNS


The strength of a round column, fig. 5.10.a, has historically been assumed
equal to that of a square member having the same area of cross-section. Therefore,
to design a round column, one may find the minimum size of square column that
would satisfy all requirements and then provide a round shape of equal area.
Tapered columns can be found in varieties, such as those shown in fig.
5.10. For tapered columns like (b) in the figure, a section at one-third of the way
from the small end is assumed to be critical. The column is designed as if the
entire column had this cross-section. However, one must also check to insure that
the compressive stress at the smallest cross-section of the tapered column does not
exceed the allowable stress.

Figure 5.10 Round and tapered


columns

(a)
108

(b)

(c)

Timber structures - 5

For columns that are not continuously tapered, such as those in fig. 5.10.c,
or for columns that increase in dimensions in one plane while decreasing in
another plane, design is based on a critical dimensions in each direction. In these
cases the dimension d in each plane of the column is taken as the sum of the
minimum d in that plane plus one-third the difference between the minimum d
and maximum d in that plane.
Round columns are often also tapered. Examples are timber piles. Round,
tapered columns are often designed as equivalent square columns, but again the
design is based on the cross-section that is one third the length from the small end.

5.7. DESIGN EXAMPLES


5.7.1. Design of an axially loaded column
The column, fig. 5.11 is subjected to an axial force coming from permanent
load, Nc1 = 150 kN and to an axial force from variable load Nc2 = 125 kN. The
column is pinned at the both ends and has a length L = 3.0 m. It is supposed that
the column has a square cross-section. The sizes of the column cross-section must
be evaluated.
Nc

Figure 5.11 Wood column


subjected to axially load
b

Design method: ASDM STAS 856-71


Wood
Geometrical properties:
Span (clear distance) L = 3000 mm;
Effective (buckling) length lf = 1.0 x L = 3000 mm.
Mechanical properties:
Species fir tree;
109

Wood column design

Moisture content 15%;


Allowable compression strength ac = 10 N/mm2 ;
Elasticity modulus E = 10000 N/mm2;
Loading:
Axially load: Nc = Nc1 + Nc2= 275 kN
Load duration class long term;
Service class 1.
Static scheme:
Double pinned column
Case 1: it is supposed
Nc
A
3100
h 1.5b

75

49899.19 [mm2]

A = 49900 [mm2] (1)

ac

A bh 1.5b 2

(2)

From (1) equals to (2) obtain:


A
182.39
1.5
h 1.5b 277.5

b = 185 [mm]
h = 280 [mm]

Now the actual slenderness ratio must be verified:

Ix

An

bh 3
338426666.7
12
bh 51800
Ix
An

ix

lf
ix

80.829

[mm4]
[mm2]
[mm]

37.11

Because the actual slenderness ratio is much smaller than the ratio supposed
at the beginning it is necessarily to change the column sizes imposed by the
condition:
75 .
h
Noted
k 1.5 we obtained:
b
110

Timber structures - 5

Nc

0.001l f

[mm2]

27503

(1)

ac

A bh 1.5b 2

h 1.5b

(2)

The condition that (1) = (2) gives:


A
135.41
1.5
h 1.5b 210

b = 140 [mm]

h = 210 [mm]

Now the actual slenderness ratio must be again verified:


Ix

An

bh 3
108045000.0
12
bh 29400

Ix
An

ix
lf
ix

[mm4]
[mm2]

60.62
49.49

The assumption of slenderness ratio is checked, therefore the sizes


bxh 140x 210 [mm x mm] are the chosen dimensions.

5.7.2. Design of a column subjected to axial force and moment


y

Load Nc

ey = 90 mm
x

250 mm

L=3.75 m

100 mm

Figure 5.12 Column details


111

Wood column design

For the design data given below, check that a 100 mm x 250 mm sawn
section, as shown in fig. 5.12, is adequate as a wall column (secondary column) if
the load is applied 90 mm eccentric to its x-x axis. The column is 3.75 m height
and has its ends restrained in position but not in direction.
Design method: LSDM NP005-96
Wood
Geometrical properties:
Cross-section dimension 100 x 250 mm;
Cross-sectional area, A = bh = 25000 mm2;
Second moment of area, Ixx = bh3/12 = 130.21 x 106 mm4;
Second moment of area, Iyy = b3h/12 = 20.83 x 106 mm4 ;
Section modulus, Wxn = 1041666.67 mm3;
Section modulus, Wyn = 416666.67 mm3;
Column length, L = 3750 mm;
Effective length, lf = L x1.00= 3750 m;
Radius of gyration, ix = 72.17 mm;
Radius of gyration, i y = 28.87 mm.
Mechanical properties:
Species fir tree;
Moisture content 15%
Characteristic bending strength, Ri = 24.0 N/mm2;
Characteristic compression strength, Rc// = 15.0 N/mm2;
Modification factor for bending:
- for permanent load, k1mod,i = mui x mdi = 1.00 x 0.55 = 0.55
Bending factor due to the preservative substances, mTi = 1.00 (no treated wood);
Partial safety factor for the material property for bending, i = 1.10;
Modification factor for compression:
- for permanent load, k1mod,c = muc x mdc = 1.00 x 0.80 = 0.80
Compression factor due to the preservative substances, mTc = 1.00;
Partial safety factor for the material property for compression, c = 1.25;
Elasticity modulus, E0.05 = 9000 N/mm2;
Working coefficient, muE = 1.00;
Factor due to the preservative substances, mTE = 1.00
Loading:
Service class 1;
Design load permanent load: 20 kN;
Axial force Nc = 20 kN;
Moment due to the eccentricity Mef,max = 1800000 Nmm.
112

Timber structures - 5

Static scheme:
Double pinned column
Check slenderness ratio:
The equations (5.3) and (5.4) are applied:

3750
72.17

51.96
max

129.89

150

3750
129.89
28.87

The maximum slenderness ratio is checked according to the permissible


slenderness ratio, table 3.6, NP 005-96.
Load capacity of the column:
CE

9000 1.00 1.00

M eff

1800000

Mr

k mod,i

1
20000
1
131440.30

Ri mTi W yn
i

3100
129.89 2
Cr

k mod,c

20830000
131440.30
3750 2

0.55

[N]

[Nmm]

2123042.921

24.0 1.0 416666.67


1.10

5000000 [Nmm]

0.184

Rc
c

Ac mTc

0.80

15.0
0.184 25000 1.0
1.25

44160

[N]

Now, the equation (5.17) must be verified:


Nc
Cr

M eff
Mr

113

Wood column design

20000
44160

2123042.921
0.87 1
5000000.00

(1)

20000
44160

2123042.921
0.0283 1
5000000.00

(2)

The equation (5.17) is satisfied. Therefore the column with the sizes shown
above will support the eccentric load.
The result obtained from equation (2) shows that there is only compression
on all cross-sectional area, and there is no change in sign of the direct stresses.

Try to
answer
these

Questions for you

1. Compare the results of the example presented in subchapter 5.7.2, which


occur in the design processes using ASDM.
2. A stud wall has an overall height of 3.0 m. The vertical studs are positioned
at 600 mm centres with spacing plates (noggins) at mid height. Carry out
design calculations to show that the studs of 50 mm x 120 mm section made
of fir-tree under service class 1 are suitable to sustain a long-term load of 8.0
kN/m.
3. Carry out the calculations using LSDM to find the column cross-section
dimensions if the column presented in task 2 is also subjected to a lateral
distributed wind load of 1.0 kN/m2.

114

Вам также может понравиться