Download as pdf or txt
Download as pdf or txt
You are on page 1of 30

I Journal of Microscopy, Vol. 104, Pr 1, M a y 1975, pp. 3-32.

Received 29 December 1974

Invited Review
Timber-a review of the structure-mechanical property
relationship

by J. M. D I N W 0 o D I E, Princes Risborough Laboratory, Princes Risborough,


Aylesbury, Bucks

SUMMARY
Following a brief introduction concerned with present consumption and future
prospects of timber and timber products the principal structure-property rela-
tionships of this low density, cellular, polymeric composite are reviewed and
discussed. Structure is examined at four levels of magnitude-macroscopic,
microscopic, sub-microscopic and chemical-and the various models used to
interpret its composite nature are described.
The dimensional instability of timber and loss of strength on wetting are
discussed in terms of its fine structure. At low levels of stressing, and for short
periods of time, timber can be treated as an elastic material, but at higher stresses
and prolonged periods, especially with alternating humidity, timber behaves as a
linear orthotropic viscoelastic material; the various factors influencing the
elastic constants and the relationship of creep to fine structure are discussed.
The anisotropy of wood is related to cell arrangement and microfibrillar
orientation: strength and its variability are discussed in terms of structure at all
four levels. Comparison of the strength of timber with that of other constructional
materials especially on a weight basis shows timber in a very good light; the
combination of high stiffness and high toughness is unique. Recent models using
three-dimensional anisotropic elastic analysis to understand strength and de-
formation of timber are described. The morphology of fracture under different
forms of stressing is illustrated.

INTRODUCTION
Those involved in research on timber would certainly agree that it is a fascinat-
ing and challenging material. This is true whether one is attempting to unravel
the mysteries still surrounding the mechanism of microfibril production and
orientation within the cell wall, or trying to explain the behaviour of this material in
terms of its complex anisotropic structure. Although a considerable amount has
been written on the physiology of wood formation, much of which is relevant to a
fuller appreciation of the behaviour of wood, it is intended that the emphasis in
this paper will be on the interrelationship of structure and properties : recourse
will only be made to the physiology of wood production to explain the formation
of different types of cells or different compositions of the cell wall.

5
J. M.Dinwoodie
The material scientist sees timber as a low density, cellular, polymeric corn-
posite and, as such, timber does not conveniently fall into any one class of material,
rather tending to overlap a number of classes. This explains in part the inadequacy
of many of the current theories evolved to explain behaviour of one class of
material to account satisfactorily for the behaviour of timber. The attempt to
explain timber performance in terms of structure is certainly not new; like other
traditional materials certain structure-property relationships were established
centuries ago. The Ancient Briton recognized the durable properties of oak
heartwood for hut and 'boat' construction, the longbowman of the middle ages
appreciated the high strength and elasticity of English yew, and the nineteenth
century wheelwright recognized the toughness of elm for the hub and the
durability of oak for the rim of his wheels.
The passage of time has resulted in more sophisticated approaches to the
establishment of structure-property relationships. Using electron microscopy or
finite element analysis, to cite examples of contrasting tools, the modern researcher
has sought to explain behaviour in terms of smaller and smaller units or even in
terms of simulated models. Whilst this is a creditable achievement and has
contributed much to our understanding of the performance of timber as a material,
a cautionary note must be recorded, for in the desire for refinement, it is easy to
overlook the significance of the grosser features. Irrespective of how significant
a relationship that can be established between tensile strength and the angle of
microfibrils in the cell wall, the presence of knots, their size and distribution, or
the slope of the grain of timber, will remain highly significant factors in controlling
the strength of timbers in practical situations.
Few readers will disagree that timber is a ubiquitous material, yet few realize
the magnitude of its consumption. On a world-wide basis 10" tonnes are used
annually, an amount almost equivalent to the annual consumption of iron and
steel (Parratt, 1972): at an average price of El50 per tonne this is equivalent to an
annual value of L150,OOO x loG.T h e United Kingdom has always been a large
consumer of timber, some 95",, of our softwood requirements and 65',,,of our
hardwood requirements coming from overseas. In 1973 the value of our imports
of timber and timber products such as various forms of board materials, but
excluding pulp, was E669 million (Anon, 1974), and, a5 such, forms about 12",,
of our total imports. Home production of timber and timber products amount to
about L50 million: of the total value of A750 million about one-third will be used
in the constructional industry. Timber is, and will continue to be used widely in
construction on account of its favourable strength-weight and strength-cost
factors.
Questions are frequently raised as to the magnitude of the timber resources in
the world and to whether the present rate of consumption can be sustained, far
less allowed to increase. The world's forests have been calculated at 3900 x 10"
hectares (FAO, 1969/70) but it has been estimated by F A 0 that from 40 to 60",,
of this forest area is unproductive. Nevertheless it has been claimed, assuming no
reafforestation or regrowth, that there is sufficient timber for several hundred
years. However, lest complacency intervene, much of this forest area is in
inaccessible parts of the world frequently containing species of wood which, at
present, qose problems in transportation and processing. T h e quantity of wood
available in the future will be very dependent on the international supply of
competing materials. Conservationists would be the first to point out that timber
is a renewable resource and the combination of regrowth in areas already felled,
together with timber from these newer areas, should meet the demand for timber
for many decades, if not centuries.

4
Timber review
Apart from its virtue as a renewable resource, conservationists also make capital
from the fact that the energy required to produce unit weight of timber is lower
than with any other material. Even after equating energy costs in terms of value
of material, timber remains considerably lower than any other structural material
(Carruthers and Hanson, 1974).

STRUCTURE
Macroscopic
The trunk of a tree has three principal functions to perform; first, it must
support the crown, a region responsible for the production not only of food but
also of seed; secondly, it must conduct the mineral solutions absorbed by the
roots upwards to the crown, and thirdly it must store manufactured food supplies
(carbohydrates) until required. As will be described in detail later, these tasks are
performed by different types of cells.
Whereas the entire cross-section of the trunk fulfils the function of support, and
increasing crown diameter is matched with increasing diameter of the trunk,
conduction and storage is restricted to the outer region of the trunk. Botanically,
this area of physiological activity is referred to as the sapwood and the corre-
sponding area in the centre of the trunk which is physiologically inactive is called
the heartwood (Fig. 1). Frequently, but certainly not always, the heartwood is
coloured due to the presence in small amounts of highly complex organic com-
pounds which confer on the heartwood some degree of durability (Hillis, 1962).
Since each species of timber has its own specific extractive(s) considerable
variation in durability will occur between timbers : sapwood is always susceptible
:o attack by insects or fungi. The width of the sapwood band decreases
n-ith increasing age of the tree and, at constant age, varies between trees of

PITH

LATE W O O D
ANNUAL OR GROWTI-
\
~ M R III M ' EARLYWOOD RING.

TWO00

RAYS

Fig. 1. Diagrammatic illustration of a wedge-shaped segment cut from


a five-year-old hardwood tree showing the principal structural features.

5
J. M. Dinwoodie
different species. Although the heartwood usually has a lower pH and lower
moisture content than the sapwood, strength of the dry timber remains constant.
Increasing size of the crown results in enlargement of existing branches and the
formation of new ones. Radial growth of the trunk must accommodate these
changes and this is achieved in the formation of the structure known as a knot. If
the cambium of the branch is still alive at the point where it fuses with the
cambium of the trunk, continuity in growth will occur even though there is a
change in cell orientation; this structure is referred to as a ‘live’ knot. If, however,
the cambium of the branch is dead, and this happens frequently on the lower
branches, discontinuity will exist and the trunk will encompass the dead branch
and its bark; such knots are termed ‘dead’ or ‘black’ and frequently drop out on
sawing. T h e grain direction around knots is frequently distorted and in a later
section the loss in timber strength associated with knot size and distribution will
be discussed.

Microscopic
The cellular structure of timber is illustrated in Figs. 2 and 3. The former is
representative of coniferous trees, known in the timber trade as softwoods,
while the latter is from a broad-leaved tree, generally known as hardwoods. In
both cases it will be observed that from 90 to 95’ ,, of the cells are aligned in the
vertical axis while the remainder are present in the form of horizontal bands
(known as rays) on the radial plane. There is no horizontal component on the
tangential plane and consequently the distribution of cells is different in each
of the three principal axes; this distribution is one of the two principal reasons for
the high degree of anisotropy present in timber.

Fig. 2. Cellular arrangement in a block of softwood (Pinrrs svluestris).

6
Timber reciew
These cells are produced by radial division of the cambium, a zone of living
cells lying between the bark (phloem) and the woody (xylem) part of the trunk
and branches (Fig. 1). With the onset of growth, the resting single layer of cells
subdivide radially to form a cambial zone some ten cells in width (Bsnnan, 1955);
the newly formed vertical wall is referred to as the primary wall. During the
growing season these cells undergo further subdivision : some of the products
lying to the inside of the cambial zone will continue to subdivide while those on
the edge of the zone will develop into bark or wood depending on their relative
position. T o accommodate the increasing diameter of the tree the cambial zone
increases circumferentially by periodic tangential division of the cambial cells.
The length of the cells in wood is closely related to the rate of tangential
division of the cambium (Bannan, 1951, 1960).

Fig. 3. Cellular arrangement in a block of hardwood (Quercm r&r).

The newly produced cells formed towards the woody tissue undergo a series of
changes extended over a period of about 3 weeks. During this process of differ-
entiation, changes in cell shape are paralleled with the formation of a secondary
wall and, towards the end of the process, by the death of the cell; the degenerated
cell contents are frequently found lining the cavity of timber cells. The newly-
formed cell is transformed into one of four basic cell types (Table 1). In Fig. 2, two
types of cells can be observed; those positioned vertically are responsible for both
the conducting and supporting r6les. Known as tracheids, these cells are from
2 to 4 mm in length with an aspect ratio of 100: 1. The horizontally aligned ray
J . M. Dinwoodie
comprises principally parenchyma cells. These cells, some 200 x 30 pm in size,
are responsible for the storage of carbohydrate. Certain softwoods possess small
amounts of vertically aligned parenchyma. In contrast, in the hardwoods (Fig. 3)
four types of cells are present, albeit that one, the tracheid is present in small
amounts. The r61e of storage is again assumed by the parenchyma which can be
present vertically as well as horizontally. Support is effected principally by long
and very narrow cells with finely tapered ends; known as fibres these cells are
usually 1-2 mm in length with an aspect ratio of 100: 1. Conduction is carried
out in cells whose end walls have been dissolved away either completely or in
part. Known as vessels these cells are short (0.2-1.2 mm) and relatively wide
( < 0-5mm) and when positioned in a vertical stack form an efficient conducting
tube. Thus, in softwoods the three basic functions of the trunk are satisfied by
two cell types while in the hardwoods each function is performed by a single type
of cell (Table 1).

Table 1. The functions and wall thickness of the various types of cells found in
softwoods and hardwoods
Cells Soft- Hord- Function Wall thickness
wood wood

Parenchyma J J Stcroge
1
Trocheids J J
Support
conduction
Jl
Fibres J Support

Vessels bores) J Conduction

The thickness of the walls of these cells is related to the task the cells perform
(Table 1). Density is a function of the ratio of wall thickness to diameter of the
cell cavity; consequently the density of wood, and to a very large extent the
strength of wood, is related to the relative proportions of the various cell types
and even more so to the wall thickness of any one type, but especially the fibres:
this has been shown to vary considerably between different species. The range in
density of timber is from 120 to 1200 kg/m3 corresponding to pore volumes of
92-1 8"".
Growth may be continuous throughout the year in certain parts of the tropics
and the timber produced is uniform in structure. Elsewhere growth is seasonal
and the timber is characterized by the presence of rings. In the early part of the
growing season the principal requirement appears to be conduction while later
in the season, support assumes dominance. This change in requirements by the
tree manifests itself in the softwoods by the presence of thin-walled tracheids
( - 2 pm) in the early part of the season, the 'early wood' or 'spring wood', and
thick-walled ( < 10 pm) slightly longer (+ lo",,) tracheids in the latter part of the
season, the 'late wood' or 'summer wood' (Fig. 2).
8
Timber review
In some hardwoods, but certainly not all, the early wood is characterized by
the presence of large diameter vessels while the late wood vessels are much
smaller and the fibre content higher. Timbers with this characteristic two-phase
system are referred to as ring-porous (Fig. 3). The majority of hardwoods,
however, show little differentiation between early wood and late wood : distribu-
tion in type and size of cells is uniform across the ring and these timbers are
referred to as diffuse-porous.
With the exception of the vessels in hardwoods, timber is basically a closed-cell
structure. Flow of liquids between cells is accommodated by the presence of pits
or openings in the cell walls. Different types of pits are to be found interconnect-
ing different combinations of cell types.
The ease with which wood can be impregnated by artificial preservatives is
related to the structure and distribution of the types of pits present (Bolton &
Petty, 1975).

Chemical and ultrastructure


A combination of chemical analyses and X-ray diffraction analysis reveals the
existence of four constituents possessing very different chemical structure (Table
2).
The extractives have already been mentioned in connection with heartwood
formation and are quite divorced from the three other structural constituents of
timber.

Table 2. Chemical composition of timber


Molecular
weight Polymeric state derivatives Function
Cellulose 45-50 Crystalline, highly Glucose ‘fibre’
orientated large
molecules
Hemicellulose 20-25 Semi-crystalline, Non-cellulosic
small molecule polysaccharides ‘matrix’
Lignin 25-30 Amorphous, large 3D Phenol propane
molecule
Extractives 0-10 Non-polymeric e.g. Terpenes poly- extraneous
phenols

Cellulose (C,H,,O,), one of the few natural polymers, occurs in the form of
long slender filaments or chains, these having been built up fairly slowly within
the cell wall from the glucose monomer. Whilst the degree of polymerization can
vary considerably from one timber to the next it is considered that there are from
8000 to 10,000 units on average (Goring & Timmell, 1962). The anhydro-glucose
unit, which is not quite flat, is in the form of a six-sided ring consisting of five
carbon atoms and one oxygen (Fig. 4); the side groups play an important r61e in
intermolecular bonding. Successive units in the chain are rotated through 180 ’
and covalent bonding is by ,B-1-4 linkages giving rise to a fairly straight and stiff
chain.
Although cellulose may crystallize in a number of forms, it is present in timber
in only one of these-cellulose I. X-ray diffraction analysis indicates that the
cellulose crystal is characterized by a repeat distance of 1.03 nm, (equivalent to
two anhydro-glucose units) in the chain direction (b-axis), with the other edges
of the unit cell having lengths of 0.835 nm in the [ 1001 crystallographic direction
(a-axis) and 0.790 nm in the [OOl] direction (c-axis): the a- and c-axes are inclined
at 84’ to each other and both perpendicular to the 6-axis (i.e. monoclinic). These
spacings and angle have been fitted into a number of unit cells over the years and
9
J . M.Dinwoodie

- - .
-
CLllobiOsC ~ n i l
- --

Fig. 4. Structural formula for the cellulose molecule.

the one which has found most acceptance is that proposed by Meyer & Misch
(1937) with the chains in a slightly bent configuration as proposed by Hermans,
De Booys & Maan (1943) (Fig. 5a); this allows H-bonding between the 3 and
5 -OH groups within the molecule and gives satisfactory interpretation of the
polarized infra-red data.

Fig. 5. (a) Side view and (b) plan view of the unit czll of cttllulostt 1
according to Liang & Marchessaut (1959). Interplane and intraplane
hydrogen bonding is indicated. (Drawing modified \lightly from that
presented by Mark, 1967). yz are H bonds in 101 plane; - - - r5
are H bonds in 101 plane; - - - x are intrxha in H bonds.

The chain at the centre of the unit is displaced longitudinally by 0.29 nm thus
facilitating inter-chain hydrogen bonding. Although there is no direct evidence
as yet that adjacent chains are lying in opposite directions there is considerable
indirect evidence to suggest this possibility (Frey-Wyssling & Muhlethaler, 1963):
Preston (1959) and Roelofsen (1959), however, dispute this hypothesis. The
possible antiparallel arrangement must not be interpreted as inferring that the
cellulose molecule is folded as is found in many of the crystalline man-made
polymers. Although a folded chain model was proposed some years ago (Manley,
1964) and recently revived (Chang, 1971) it can satisfy only a certain number of
the physical observations and the model is now disputed in favour of extended
chain structures (Mark et al., 1969; Frey-Wyssling & Muhlethaler, 1965).
Within the structureboth primary and secondary bonding is represented: covalent
bonding both within and between the glucose rings manifests itself in the high axial
tensile strength of timber. Lateral bonding appears to be limited to a mixture of
hydrogen bonds and van der Waals forces. Although Meyer & Misch (1937) placed
the hydrogen bonds within the (002) plane, it is now thought that these bonds
unite cellulose chains in the (101) and (lO'i) planes (Fig. 5b) thereby providing
10
Timber review
the principal mechanism for stabilizing the crystal against relative displacement
of the chain and contributing considerably to the axial stiffness of timber (Jaswon,
Gillis & Marks, 1968). The same O H groups that give rise to this hydrogen
bonding are highly attractive to water molecules and explain the affinityof cellulose
for water.
As Table 2 indicates, timber is a collection of crystalline and non-crystalline
substances: the degree of crystallinity for timber as a whole has been shown to
be at least 67O, and not greater than goo,. Regions of complete crystallinity give
way gradually to regions totally devoid of crystalline structure. The length of the
cellulose molecule is about 5000nm, whereas the length of a crystalline zone
(crystallite) is only about 60 nm: consequently any one cellulose molecule will pass
through several crystallites with intermediate regions lacking in or possessing
only slight crystallization.
The model which best fits experimental results is the fringed fibril concept
proposed by Hearle (1963). Although certain chains digress on emerging from a
crystallite and twist and turn before conforming to the ordered structure of a
distant crystallite, the majority of chains remain closely associated with each other.
There is a high degree of longitudinal co-ordination between the crystallites ; this
collective unit is called the microfibril. Of infinite length this fundamental unit
is from 10 to 30 nm in width.
The hemicelluloses (Table 2), like cellulose itself, are carbohydrates differing
in composition between hardwoods and softwoods. However, the degrees of
crystallization and polymerization are but a small fraction of that occurring in
cellulose.
Lignin, present in roughly equal proportions to the hemicelluloses, is chemically
dissimilar to these and to cellulose (Table 2). It is non-crystalline, differing in
structure between hardwoods and softwoods, and is deposited within the cell wall
following the completion of carbohydrate formation. Most cellulosic plants do not
contain lignin and it is the inclusion of this component which imparts much of
the stiffness to timber.
In the introductory remarks wood was defined as a natural composite and two
models analogous to coarse and fine scales, have been proposed to interpret the
ultrastructure of wood from the various chemical, X-ray and electron diffraction
analyses. The coarse model ascribes the r61e of ‘fibre’ to the actual cell and the
r6le of ‘matrix’ to the lignin-pectin intercellular layer which cements together the
cells. Whilst the model satisfies certain aspects of timber behaviour it is far from
adequate, completely failing to describe swelling of the cell wall in the presence
of moisture. In addition, the model has to consider hollow fibres, unlike the man-
made composites.
The more successful model simulates composite structure at a lower order of
magnitude, the r81e of ‘fibre’ is ascribed to the cellulosic microfibrils while the
lignin and hemicelluloses are considered as separate components of the ‘ matrix’
(Frey-Wyssling, 1968). The cellulosic microfibril is regarded therefore as con-
ferring high tensile strength to the composite and its supreme performance in
this respect is related to the presence of covalent bonding both within and between
the anhydro-glucose units. It has been shown that reduction in chain length by
gamma irradiation markedly reduces the tensile strength of timber (Ifju, 1964):
the significance of chain length in determining strength has been confirmed in
studies of wood with inherently low degrees of polymerization (Dinwoodie, 1965).
While Ifju assumes slippage between chains of cellulose to be an important
contributor to the development of ultimate tensile strength, this viewpoint is
disputed by Cowdrey & Preston (1966) and Mark (1967), calculations from both

11
J. M. Dinwoodie
papers indicating the unlikely occurrence of slippage due to the forces involved in
fracturing large numbers of hydrogen bonds. Preston (1964) has shown that the
hemicelluloses are frequently associated intimately with the cellulose effectively
binding the cellulose chains together. Bundles of cellulose chains within the micro-
fibril are regarded as having a polycrystalline sheath of hemicellulose material
(Dennis & Preston, 1961) and as a consequence of the degree of hydrogen bonding
involved it seems unlikely that slippage between microfibrils occurs. Meyer (1950)
would confirm that stressing would result in preferential breakage of a cellulose
chain at the C-0-C linkage.
The deposition of lignin varies in extent in different parts of the cell wall, but
it is now evident that its prime function is to protect the hydrophilic cellulose
and hemicelluloses which are mechanically weak when wet. Experimentally, it has
been demonstrated that removal of the lignin markedly reduces the strength of
wood in the wet state, though increases it when the wood is dry, calculated on
a net cell-wall area basis (Klauditz, 1952). Since the lignin is located only on the
exterior it must be responsible for cementing together the microfibrils thereby
imparting shear resistance in the transference of stress throughout the composite.
Some controversy exists over the detail of the model regarding the distribution
of hemicellulose and lignin. The majority of workers (e.g. Frey-Wyssling,
Miihlethaler & Muggli, 1966; Sullivan, 1968; Hcyn, 1969) agree that cellulosic
subunits, some 3 nm in diameter, known as elementary fibrils or protofibrils and
containing some forty chains, are separated by gaps 1 nm in width containing
hemicellulose. Surrounding a number of these subunits is a sheath comprising
hemicelluloses and lignin. However, this sub-division of the microfibril is con-
sidered to be due to artefacts in sample preparation for electron microscopy
(Nieduszynski & Preston, 1970; Preston, 1971) though agreement exists on the
external diameter of the microfibril (10-30 nm) and the presence of lignin only on
the outer layer. The degree of crystallinity, therefore, decreases progressively
from core to perimeter of the microfibril.
Early investigations using optical microscopy (Bailey & Vestal, 1937) and
polarization microscopy (Preston, 1947; Preston & Wardrop, 1919) indicated that
the secondary cell wall could be sub-divided into three distinct layers according
to the angle at which the microfibrils were lying. More recent work using trans-
mission electron microscopy has confirmed this (Liese, 1963; Wardrop, 1964;
Harada, 1965; Chtk, 1967). Whereas in the primary wall the microfibrils are
loosely packed and interweave at random with no indication of lamellation, in
the secondary wall formed some 3 weeks later the microfibrils are closely packed
and lie parallel to one another (Fig. 6).
The outer layer of this wall, known as the S , layer and comprising up to 5",, of
the wall thickness, is characterized by having from four to six lamellae, the
microfibrils of each alternating between a left- and right-hand spiral, both
with a pitch to the longitudinal axis of from 50- to 70 ' depending on species of
timber.
The middle layer of the secondary wall, S,, accounts for up to 85",,of the wall
thickness and comprises from thirty to 150 lamellae, all exhibiting microfibrils
spiralling in a right-hand direction with a pitch of 10' to 30-to the longitudinal
axis. I n this layer the microfibrils show a higher degree of parallelism than in any
other layer. Due to the relative thickness of this layer its ultrastructure has a
profound influence on many timber properties as will be discussed later, e.g.
anisotropy, strength, shrinkage, mode of failure.
The inner layer, the S3, which can be absent in certain timbers, is characterized
as in the S, layer by being thin (loo) and with alternating lamellae possessing

12
Timber review
Secondary wall

Mlddle layer (S2J

Middle lamella

Fig. 6. Simplified structure of the cell wall showing orientation of the


microfibrils in each of the major wall layers.

microfibrils orientated in opposite spirals with a pitch from 60 to 90 (Wardrop,


1964).
More recent work (Dunning, 1969; Harada, 1965) has indicated that systematic
variation in angle occurs within each layer and also along the length of the cell.
Local deformations in microfibril orientation occur at the openings to pits in the
cell wall. Microfibrillar angle has been shown to vary between radial and tangential
walls (Mark, 1967; Tang, 1972).
This system of filament winding to produce a rigid and tough cylinder that will
withstand high tensile and compressive stresses in the wet condition has been
evolved over millions of years. Man can only marvel at the complexity and
efficiency of the structure and has only recently been able to copy the principles
for use with synthetic and more expensive materials. He is now able to produce
materials stronger and tougher than wood but only at the expense of very much
higher energy consumptions.

VARIABILITY
Variability in timber is one of its characteristic deficiencies as a material. Differ-
ences in structure and hence performance occur not only between different species
of timber but also between trees of the same species growing in different environ-
ments or between different parts of a single tree. Within a tree trunk there are
systematic patterns of variation in cell length, wall thickness, angle of the grain
and angle of the microfibrils in the S, layer, resulting in the formation of a core of

13
J. M. Dinwoodie
wood usually about ten rings in width with many undesirable properties, including
low strength and high shrinkage. Another source of variation is the formation of
abnormal or 'reaction' wood produced in leaning or exposed stems.

PROPERTIES A N D BEHAVIOUR OF TIMBER


Water relationships
Timber in the green state, i.e. newly felled, has moisture contents ranging from
60 to 200",,based on the oven dry weight of the timber. Timber can be used at
these moisture contents only where it is immersed in water, e.g. harbour piling:
for all other uses the moisture content must be lowered to provide dimensional
stability and to reduce the risk of attack by fungi. Reduction in moisture content
can be effected over a long period of time by stacking in an open-sided shed, or in
a relatively short time in a drying kiln where the combined action of heat and
high humidity is used to dry the timber at a controlled rate, thereby reducing the
effects of degrade through splitting and twisting.
At moisture contents from 200",, down to about 27'j:,, water is present both
within the cell wall and within the cell cavity. Less energy is required to remove
the latter and on drying down to 27",, no dimensional changes occur and there is
no effect on timber strength (Fig. 7). Below 27",,, a somewhat arbitrary value

0
I I
20
I
40
I
60
I
80
I
I00
Molslure content i%l

Fig. 7. Relationship of longitudinal comprcssive strength to moisture


content of the timber.

known as the 'fibre saturation point', water is present only within the cell wall:
consequently on its removal from between the microfibrils (and X-ray diffraction
analysis indicates that the water does not enter the core of the microfibril) these
will move into closer proximity because of hydrogen bonding and the timber will
shrink. Strength will correspondingly increase and Fig. 7 illustrates a three-fold
increase in strength on reduction of moisture content from 25",, to zero, in
samples of Scots pine timber.
14
Timber review
Shrinkage, however, is not uniform in all directions and Table 3 provides some
idea of the degree of anisotropy for three timbers covering the complete range of
values.
The very marked difference in shrinkage between the longitudinal value and the
two horizontal ones can be related to the angle of the microfibrils in the S, layer.
Since these are inclined at an angle of 10-30 ’ to the vertical, the horizontal com-
ponent of the movement nearer each other will be considerably greater than the
longitudinal component. A non-linear relationship between degree of shrinkage
and microfibrillar angle has been established (Kelsey, 1963; Harris & Meylan,
1965; Sadoh & Kingston, 1967; Meylan, 1972). Fairlysuccessful models for timber
behaviour based on fibre composite theories have been proposed (Barber &
Meylan, 1964; Barber, 1968; Barrett, Schniewind & Taylor, 1972; Cave, 1972)
and the relationship of swelling to laminar sorption has been studied by Christenen
(1967) and Stamm & Smith (1969).

Table 3. Shrinkage on drying from green to 12”,,moisture content


Shrinkage (‘I<,)

Tangential Radial Longitudinal


Yellow pine 3.5 1.5 0.1
Scots pine 4.5 3.0 0.1
Beech 9.5 4.5 0.75

Differences in horizontal shrinkage between the tangential and radial planes


have been explained in terms of the presence of the rays on the radial plane
producing a restricting effect, and the increased thickness of the middle lamella
on the tangential plane compared with the radial plane.
Since the moisture content of timber will eventually come to equilibrium with
the relative humidity of its environment with corresponding changes in its
dimensions, it is essential that timber should be dried to the appropriate level
prior to usage. Thus, while exterior joinery timber need be dried only to 19”,,
moisture content, timber to be used in centrally-heated rooms must be dried to
lo”,, moisture content prior to use. Fluctuations in environmental humidity will
result in small ‘movement’ of wood for the same reasons as the occurrence of
shrinkage. In addition to the effect of moisture content on strength (Fig. 7)
increasing moisture content of timber decreases stiffness and toughness, ‘green’
ash being an exception to this general rule.

Stress-strain relationships
In an idealized tree trunk, timber exhibits circular symmetry about the core
and its anisotropic elastic properties also possess circular symmetry. Conse-
quently in any large piece of timber which includes the centre of the tree, a uni-
form stress will not produce a uniform strain. However, samples of timber which
are small in comparison to the distance from the tree centre approach the ideal
state where the three principal directions of the grain lie along three mutually
perpendicular axes. In reality, the radial axes in timber are diverging, and the
longitudinal-tangential surface is not planar but cylindrical ; however, by con-
sidering small samples of timber and assuming it to be a homogeneous material
(see below) with three mutually perpendicular axes of symmetry, an elastic theory
for timber behaviour can be formulated by treating the system mathematically as
rhombic (Barkas, 1946; Hearmon, 1948). Certainly under small strains it can be
assumed to follow Hooke’s law.
The assumption for the purpose of analysis that wood is homogeneous appears
15
J . M. Dinwoodie
to be valid. Although the presence of cells makes the material inhomogeneous on a
microscopic scale Price (1928) has shown theoretically that, in units large com-
pared with the dimensions of the cells, or with the width of the annual rings, the
gross structure may be treated mathematically as a homogeneous material.
Elasticity theory describes the behaviour of a homogeneous material in terms of
thirty-six constants which, on account of certain mathematical relationships, can
be reduced to twenty-one. The treatment of timber as a rhombic solid with
inbuilt symmetry further reduces the number to twelve. Three values of Poisson’s
ratio can be calculated leaving nine independent constants to describe the
behaviour of wood when stressed parallel to the principal axes. The values of
these three Young’s moduli, three rigidity moduli and three Poisson’s ratios for a
number of timbers are presented in Table 4; this provides an indication of the
values of these constants in timber and the degree of variation that occurs among
different timbers.
Table 4 indicates that the modulus of elasticity along the grain (El,)is consider-
ably greater than that across the grain (&, &): the degree of anisotropy varying
from 18 to 60. Values as high as 182 have been recorded for certain timbers
(Kollman & CBti, 1968); generally, softwoods have a higher degree of anisotropy
than hardwoods due probably to the lower variability in cellular structure in the
former. In comparing materials, timber is characterized by having the highest
degree of anisotropy: even when the comparison is with materials having the
same elastic constants ( e g crystals) the degree of anisotropy is considerably
greater, e.g. 18-182 compared with a maximum of 2 for most crystals (Hearmon,
1953).
The degree of anisotropy in both stiffness and shear modulus (Table 4)
between longitudinal and horizontal planes in timber is due in part to the longi-
tudinal arrangement of the individual cells and in part to the orientation of the
microfibrils in the S2 layer of the cell wall (Cowdrey & Preston, 1966; Cave, 1968).
Stiffness in the radial horizontal plane appears to be about 50”,, greater than in
the corresponding tangential plane due, in some considerable degree, to the
presence of horizontally aligned cells on the former plane.
An important property of anisotropic materials is the change in values of the
stiffnesses and compliances with change in orientation of applied stress. The
effects of rotation on the elastic constants can be obtained by transforming the
stresses or strains from the rotated to the principal axes since the potential
energy of a deformed body is independent of orientation. The derived equations
(Hearmon, 1953; Kollman & C6tC, 1968) indicate that solutions of problems in
orthotropic elasticity are likely to be complicated and until the arrival of com-
puters, few three-dimensional problems were attempted : the simplified two-
dimensional approach has been used in the mathematical simulation of a matrix
of non-crystalline material (Cowdrey & Preston, 1966) or as a layered system
(Mark, 1967). Mark’s single fibre model treated the fibre as slit end to end and
opened out as a two-dimensional shb. Shear restraint was considered only in the
S, layer.
Cave (1968) was the first to treat the cell three-dimensionally, albeit in a simple
form. He obtained the elastic constants of the wall in terms of the elastic constants
of a set of parallel fibres embedded in an isotropic matrix and the angular distribu-
tion of the fibres and he represented the restriction of rotation of the cell by its
neighbours in mathematical terms by a crossed fibrillar structure simulating the
reversed helices in adjacent cell walls. The use of this two-ply balanced-layer
concept simplified considerably the elastic analysis.
Cave’s concept of complete shear restraint within the cell wall was taken up by
16
Table 4. Elastic constants of woods (1: and G in N mmi)
Specjfic Moisture
Species gravity content El, Ell 15,. r [*I I< !*I 'r GI T GI 11 G, 1%
(I'o)

Ash 0.67 9 15790 1516 a27 4895 3172 3516 896 1310 269
Balsa 0.20 9 6274 296 103 4551 1586 3379 200 310 33
Walnut 0.59 11 11239 1172 62 1 4964 3379 4344 690 896 228
Scots pine 0.55 10 16272 1103 573 4689 2896 3516 676 1172 66
Sitka spruce 0.39 12 11583 896 496 2965 2551 3241 690 758 39
El,, Modulus of elasticity in longitudinal direction. E,, Modulus of elasticity in horizontal radial direction. ET,Modulus of elasticity in horizontal
tangential direction. GI,,, Shear modulus in longitudinal tangential plane. GI,,, Shear modulus in longitudinal radial plane. G,,, Shear modulus in hori-
zontal plane. pI1~,Poisson's ratio for extensional stress in radial direction = compressive strain in T directioniextension strain in R direction. pI.R,
Poisson's ratio for extensional stress in longitudinal direction = compressive strain in R direction:extensional strain in L direction. pI.T,Poisson's ratio
or extensional stress in longitudinal direction = compressive strain in T direction kxrcnsional strain in L direction.
J. M. Dinwoodie
Schniewind & Barrett (1969) whose three-dimensional model is similar to that of
Cave's but now takes into account the layered structure of the cell wall. Mark &
Gillis (1970) in reviewing these models and in carrying out both physical and
mathematical modelling consider that complete shear restraint is virtually un-
obtainable: on the other hand, restraint only in the S, (Mark, 1967) allows more
shear strain than occurs in timber, and the former model is considered to be nearer
to physical reality. They developed the model to permit the tangential and radial
walls to undergo different strains.
Tang (1972) in attempting to derive the distribution of stress in each layer of
the cell wall when the cell is deformed, modelled his cell as a hollow, composite,
anisotropic cylinder possessing two different sets of elastic constants for the
radial and tangential walls. The results of his three-dimensional analysis provided
a completely different distribution of stresses than recorded with the two-
dimensional models and also gave information on the relative twisting angle of
each layer of the cell wall and of the fibre as a whole. Subsequently, refinements
were made to the model and to the analysis (Tang & Hsu, 1973). However, Barrett
& Schniewind (1973) are critical of some of the assumptions made in Tang's
analysis, especially one of his boundary conditions, and using three-dimensional
finite element analysis have presented a solution for a concentric, multilayered
orthotropic cylinder with uniform loading. Good agreement arose between their
results and the earlier ones from two-dimensional analysis with maximum axial
forces differing by less than 3",,.
Although the effect of direction of applied stress relative to the fibre axis on the
stiffness of timber is considerable, little difference occurs among the different
modes of stressing. The moduli of elasticity in tension, compression and bending
are all approximately equal for small stresses though the elastic limit is con-
siderably lower for compression (30-50",, of the ultimate stress) than for tension
(60-70';,,). A deflection curve for tensile stressing of beech is presented in
Fig. 8: only slight departure from linearity above the elastic limit occurs. For
small levels of stress and short period of time, given constant conditions of
humidity and temperature, timber can be considered an elastic material.
Elastic constants vary with density, moisture content, temperature and time :
in many instances the variables are interrelated. A linear positive relationship
has been established between the modulus of elasticity and density for a range of
timbers including spruce, oak (Kollmann & Krech, 1960) and balsa (Draffin &
Miihlenbruch, 1937); such a relationship is to be expected since density is but a
function of the ratio of cell wall thickness to cell diameter, and variation in either
parameter will markedly affect the stiffness of the cylindrical cell.
The location of moisture within the cell wall has already been described and
Fig. 7 illustrated the marked increase in strength on reducing moisture content
below the fibre saturation point. A similar pattern has been established for the
relationship between stiffness and moisture content with increases in modulus o f
about 50°,, on reducing moisture content from 25",, to zero (Thunell, 1941;
Hearmon, 1953; Kollmann & Krech, 1960).
I n common with many other materials increasing temperature has been shown
to decrease the stiffness of wood (Thunell, 1941; Kollmann & CGte, 1968).
Sulzberger (1948, 1953) carried out a comprehensive series of tests on a wide
range of timbers and in addition to confirming the inverse relationship of stiffness
and temperature, illustrated that the relationship is moisture dependent, sen-
sitivity increasing with increasing moisture content. Some of his results are
presented in Table 5 from which it will be observed that the effect of temperature
is more critical with increasing moisture content. It should be realized, however,

18
Timber review
that in practical terms, increasing temperature will result in a drying out of the
timber and raising of the stiffness; this may well offset the reduction in stiffness
due to increasing temperature per se (Hearmon, 1953).
Since wood is anisotropic and knots are characterized by the occurrence of
distorted grain, it follows that the presence of knots will reduce stiffness, this
reduction being related to the size, location and frequency of the knots.

3300 - 2
,

I //”
I.
7500 - ld
/I
I,.

4’
2300 -
4‘
I

t 1500- r’
1

I000 -
r‘
t’
r’
500-
{‘
Fig. 8. Load extension diagram for dry beech wood under tensile
stressing.

Although it is possible to consider timber as an elastic material where the


level of stress is low and is effective for instantaneous time, timber, like many
other high poiymers, should really be treated as a viscoelastic solid. Under con-
ditions of prolonged moderate to high stressing or under lower stress levels with
cyclic changes in humidity, timber has been shown to exhibit considerable plastic
flow. Initial loading of a sample followed by release of the load will result in
permanent set: if the sample is now loaded in reverse and subsequently released
a hysteresis loop of stress-strain behaviour is formed, the area within which
represents the energy loss in the cycle of stressing (Barkas, 1949). When the
stress is applied in a single direction and repeated at regular intervals, the total
plastic flow or permanent set increases, though the increase in set per cycle
actually decreases (Barkas, 1949).
T h e stress level at which set first begins to appear varies with species and testing

Table 5. The interdependence of temperature and moisture content on the


stiffness of wood
Moisture Relative stiffness where E,, = loo1,’(,
content Temperature -C
(“J 60 40 20 0 - 20
0 97 98 100 102 104
8 89 96 100 103 106
12 84 93 100 104 108
20 73 89 100 107 113

19
J . M.Dinwoodie
conditions, but has been shown to range from 21 to 80",, of the ultimate tensile
strength with an average about 40",, (King, 1961). Recent work by Schniewind &
Barrett (1972) has confirmed that wood can be treated as a linear orthotropic
viscoelastic material, at least to a first approximation, and demonstrated somewhat
surprisingly that there is an increase in volume when wood is stressed in uniaxial
tension parallel to the grain.
The application of stress to a viscoelastic material such as timber results
initially in instantaneous elastic deformation, followed by a period of retarded
deformation under constant load, known as creep. When stress is removed, part
of the deformation instantaneously disappears (elastic recovery) followed by a
period of retarded partial recovery known as primary creep; the non-recoverable
portion is termed secondary creep. T h e elastic or primary creep is due in a
considerable extent to the change in moisture content in wood under stress.
Timber under conditions of constant humidity will lose water if subjected to
compression and gain water under tensile stressing: in both cases there is a time
delay due to the slow rate of diffusion of water out of or into the wood (Barkas,
1946).
The amount of secondary creep is proportional to the logarithm of time in the
initial stages but increases more rapidly beyond this period (King, 1961). Time to
failure will obviously increase as the level of applied stress is lowered, while
increased deflection or creep rate will rise at higher moisture contents (Clouser,
1959) and elevated temperatures (Davidson, 1962; Kingston & Budgen, 1972).
Creep compliance also increases with stress, but the relationship is linear only
to about 60"; of ultimate stress in bending and from 60-70",, in compression,
above which levels compliance increases rapidly (Kingston & Clarke, 1961;
Kingston & Budgen, 1972; Keith, 1972). A similar departure from linearity has
been noted in stress relaxation in tension and compression, stress relaxation being
a non-linear function of strain at all levels of strain (Echenique-Manrique, 1967).
Viscoelastic behaviour in timber has been interpreted using a series of models
combining springs and dash pots to simulate elastic and plastic regions. At a
microscopic level Kollmann (1961) used a series of these models to demonstrate
the effect of vessel size, ray width and relative proportion of early and late wood
on the mechanical properties of timber. However, the use of this type of model
has generally been applied to study the behaviour of the material at a sub-
microscopic level where the springs simulate the highly crystalline elastic regions
and the dash pots the amorphous zones. Allowing for the fact that in timber these
zones are never isolated but fuse into each other, remarkably good success has
been achieved in describing creep behaviour in terms of these models. Senft &
Suddarth (1971) using only three and four element models were able to explain
80°,,and 94q4 respectively of the variation in creep response in Sitka spruce
timber. Modelling has also been used in relaxation studies (Echenique-Manrique,
1967).
Attempts have been made to describe creep in timber in terms of its fine
structure. El-osta & Wellwood (1972a, b) have shown that short-term creep
was highly correlated (f0.82) with the angle of the microfibrils in the S,
layer of the cell wall (1972a) and also correlated inversely with the degree of
crystallinity (1972b).
The possible r81e of hydrogen bonds in creep response is demonstrated by the
behaviour of wood under load during a change in moisture content. It has been
observed under cyclic changes in moisture content (Armstrong & Christensen,
1961; Hearmon & Paton, 1964) that although the deflection behaves in a cyclic
manner, the recovery in each cycle is only partial and hence total creep is large:

20
Timber review
rhe modulus of elasticity remains unchanged. The evidence suggests that creep
xkes place during the movement of moisture through the wood, probably in a
series of steps from one absorption site to another, involving a breaking of
iydrogen bonds and temporary weakening of the timber (Gibson, 1965). Nissan
h Sternstein (1961) also explain viscous behaviour in terms of time-dependent
change in the active number of hydrogen bonds. More recently, rheological
t.shaviour in timber has been explained in terms of time-dependent, two-stage,
molecular motions (Chow, 1973).
Readers anxious to study timber rheology are referred to a comprehensive
review of the subject by Schniewind (1968).

Srrength
Over the years a number of models have been employed in an attempt to
quantify the theoretical tensile strength of timber. In these models it is assumed
that the lignin and hemicelluloses make no contribution to the strength of the
timber. One of the earliest attempts regarded timber as comprising a series of
endless chain molecules and strengths of 8000 N/mm2 were calculated (Meyer &
.!lark, 1930). Recent modelling has taken into account the finite length of the
cellulose molecules and the presence of amorphous regions. The calculated stress
to cause chain slippage is generally greater than that to cause chain scission
irrespective of whether the latter is calculated on the basis of potential energy
function or intrachain linkbond energies. The minimum derived stresses are of
the order of 1000-7000 N/mm2 (H. Mark, 1952, R. Mark 1967).
The ultimate tensile strength of timber is of the order of 100 N/mm2, a
value corresponding first to between one tenth and one eightieth of the theoretical
strength of the cellulose fraction, and secondly, to about one hundredth the
modulus of elasticity of timber. Strain at failure in timber is about lo, compar-
able with that of carbon fibre reinforced plastics.
It is possible to separate the individual cells by dissolution of the lignin-pectin
complex cementing them together and tensile stressing of these has indicated
strengths in the order of 500 N/mm2 equivalent to half the theoretical value
(Dinwoodie, 1965; McIntosh, 1965).
Since timber is seldom stressed in pure tension in industrial usage, interest in
its tensile strength remains somewhat academic. Since this property represents
one of the principal attributes of this material it is unfortunate that so little use
is made of this feature. The most frequent application of stress is in a bending
mode whilst its resistance to compressive loading in an axial plane is also of
considerable practical significance.
Timber is one of the few materials which possesses high toughness as well as high
stiffness and this unique combination of properties accounts for many of the
specialized applications of timber. The order of magnitude of these mechanical
strength properties, together with an indication of the degree of interspecific
difference that occurs, is indicated in Table 6. The timbers are separated into
softwoods and hardwoods as defined in Table 1 and it will be observed that these
terms bear no relation to the density of the timber; the range in density of the
hardwoods overlaps that of the softwoods.
Bending strength is approximately one-half the tensile strength along the grain
while resistance to longitudinal compression is about one-quarter of the
corresponding tensile strength. Toughness in timber is generally measured as the
resistance to impact and is usually determined on a modified Hatt-Turner mach-
ine. In the table, stiffness is included for comparative purposes: the entire
range of tests performed on timber, together with test methods and data on a very
21
J . M . Dinwoodie
Table 6. Strength and stiffness of certain timbers
Bending Compression Impact
Density (modulus) (maximum (maximum E
of rupture) strength drop of
paralell to hammer)
grain)
Softwoods (kg,m’) (N mm2) (N mm’) (m) (Nmrn2)
Western red cedar 370 65 35.0 0.58 7,000
Scots pine 510 89 47.4 0.71 10,000
European larch 560 92 46.7 0.56 9,900

Hardwoods
Balsa 180 23 15.5 3,200
Obeche 350 54 28.2 0.48 5,500
Beech 690 108 51.8 0.99 10,100
Greenheart 990 181 89.9 1.35 2 1,000

comprehensive range of timbers, is presented in Bulletin 50 of the Forest Products


Research Laboratory (now the Princes Risborough Laboratory) (Lavers, 1969).
Table 6 illustrates the high degree of correlation that exists between density
and strength. Very high correlations are obtained with compression strength and
hardness, while lower values are recorded for tensile strengths and cleavage.
Density, as recorded earlier, is a reflection of the ratio of cell-wall thickness to cell
diameter and the interspecific differences in strength can be related to the
different proportions of cell types (Table 1) or differing thickness of cell wall of
any one type; these parameters are influenced considerably by the rate of growth
of the tree.
I n previous sections timber was shown to be anisotropic in terms of stiffness
and shrinkage. For exactly the same anatomical reasons timber is also anisotropic
with regard to strength; values for Douglas fir are presented in Table i .
Irrespective of the moisture content of the timber the highest degree of
anisotropy is in tension. When timber is stressed in tension along the grain,
strength is higher than for any other mode; correspondingly, when stressed in
tension across the grain the strengths recorded are the lowest for all modes of
stressing. A similar degree of anisotropy is present in tensile stressing of both
glass reinforced plastic and carbon fibre reinforced plastic when the fibre is laid
up in parallel strands.
Table 7 also illustrates that the degree of anisotropy in compression is about
an order of magnitude less than in tension.
Since timber is an anisotropic material it follows that the angle at which stress
is applied relative to the longitudinal axis of the cells will determine the strength
to failure. Strength in tension appears to be much more sensitive to angle of the
grain than either bending or compression and at angles exceeding 8 there is a
marked decline in tensile strength.
The influence of moisture on strength has already been noted especially the
very marked increase in strength with reduction in moisture content below 25(’,,
(Fig. 7). Table 7, whilst reinforcing this relationship, also illustrates that stressing
in tension is much less sensitive to level of moisture than is compression stressing.
In common with many other materials timber displays an inverse relationship
between strength and temperature (Lavers, 1969). Although the effect of tempera-
ture on tensile strength appears to be slight (Kollmann & C6t6, 1968) this is
certainly not the case with other modes of stressing and in bending and toughness
the effect of temperature is considerable. In the case of modulus of rupture,
increasing temperature will have a similar effect to increasing moisture content;
Sulzberger (1948) has indicated a linear relationship within the ranges of 6-20qb
22
Timber review
Table 7. Anisotropy in strength in Douglas fir timber
Moisture Tension Compression
Content ;I - // + I /i - - -

("u) N/mm* Nzmrn' N!mm* N,'mm*


> 25 131 2.69 48.7 24.1 4.14 5.82
12 138 2.90 47.6 49.6 6.90 7.19

moisture content and - 20 C to 60 C, thereby allowing transformation of results.


However, in the case of toughness, moisture content and temperature interact
rather than complement; at low moisture content strength increases with de-
creasing temperature, while at high moisture content the reverse relationship
occurs (Thunell, 1941).
Not only are the size and number of knots important in determining the degree
of strength reduction, but the type and location of the knots are also critical
factors. In general terms 'live' knots confer lower strength losses than do 'dead'
knots, while knots on the edge of a piece of timber are more critical than in the
centre, especially so when subjected to bending stresses. Tensile stressing is
particularly sensitive to the presence of knots.
The effect of knot size can be evaluated using the 'knot ratio' which is the
ratio of knot diameter to the width of timber section. Sunley (1968) found a linear
negative relationship between strength and knot ratio for small and medium-sized
knots. Knots with diameters up to one half the width of timber section will result
in strength reductions of 50°,,.
Recently fracture mechanics has been successfully applied to the study of
tensile strength of structural timber and the relative effects of a whole series of
defects including knots were evaluated using this tool (Schniewind & Lyon, 1973;
Pearson, 1974). In addition to knots and splits, strength can be influenced
considerably by such defects as reaction wood and the presence of compression
damage formed as a result of growth stresses within the tree (brittleheart) or
subsequent damage during felling or handling. Reductions of up to 50",, in
toughness have been found in material containing compression damage (Din-
woodie, 1975b).
Various attempts have been made to relate strength, especially tensile strength,
to the angle of the microfibrils in the various cell wall layers (Mark, 1967;
Schniewind & Barrett, 1969; Mark & Gillis, 1970; Tang & Hsu, 1973)-many
of these investigations were also concerned with relating stiffness to micro-
fibrillar orientation and have already been discussed.
I n an assessment of the various inherent and environmental factors influencing
strength the effect of time must not be overlooked. The effect of duration of load is
inversely related to degree of stressing (Clousser, 1959; Pearson, 1972) and
timber subjected to constant stress for long periods of time will fail a t stresses
considerably lower than those it can withstand for short durations. It is considered
that after 50 years of constant stress the strength of the timber will be rcduced to
about 50°, of the original. In Pearson's review (1972) he established a linear
relationship between strength loss and time finding no evidence for the existence
of a threshold stress below which failure would not occur; if such does exist he
claimed that it must be less than 50"" of the short-term strength. Certainly earlier
work (Wood, 1951) has shown a negative hyperbolic relationship suggesting a
flattening out between 40 and 50°, of the short-term load. In the calculation of
working stresses for timber, allowances have to be made in the safety factor to
cover not only snow and wind loads but also duration of the dead load: the
safety factor adopted in the UK for bending, tension and shear stresses is 2.25
(Sunley, 1968).

23
J. M . Dinwoodie
Timber has been used as a construction material for hundreds of centuries and
in many of its strength properties it compares favourably with other constructional
materials especially so when comparisons are made on a strength-weight basis
(Table 8).

Table 8. Strength and stiffness of wood (constructional 'whitewood') in


comparison with other materials
Specific Com-
Material Sp.Gr Tensile tensile E E,Sp.gr pression
(MN m') (MN m-') (GN/m2) (GN m') (MN m')
Wood (spruce) 0.46 104 226 10 22 37
Concrete 2.5 4 3
I 40
~~ 19
- . 69
Glass 2.5 50 20 69 28 50
Aluminium (Duralumin) 2.8 247 88 69 25 -
Cast iron 7.8 138 18 207 26 120
Steel (mild) (0.06",, 7.9 459 58 203 26 800
carbon)
ABS 1.1 50 45 3 3 50
PVC (rigid) 1.5 59 39 2.4 1.7 55
Polyester resin G cloth 1.8 276 153 18 10 270
- - unidirectional
Epoxy 1.8 1100 61 1 45 25 400
roving
CFRP 1.5 1040 693 180 120 1040

It will be observed that the tensile strength of constructional grade softwood


is in the middle of the range of quoted results; however, due to the low density
of this cellular material its tensile strength per unit weight (specific tensile
strength) is well ahead of most traditional constructional materials. The use of
timber for the construction of the Mosquito aircraft during the war and its
continued use for glider production bears testament to its high strength-to-weight
ratio.
While the stiffness of wood is lower than that of concrete and the metals in
absolute terms, it assumes equality when expressed in terms of weight; only carbon
fibre reinforced plastic (CFRP) has a specific stiffness markedly superior to that
of timber.
Timber continues to be a low-cost material with a price for constructional grades
of about 13p per kg (1974 prices). When strength is related to cost, timber is
found to compare very favourably with most other constructional materials.
Timber is a tough material, its fracture energy or work of fracture, while being
an order of magnitude less than that of ductile metals, is comparable with that of
the man-made composites (Table 9) (Gordon & Jeronimidis, 1974). It is generally
considered that for structural applications a material should possess a work of
fracture in the region of 104-106J,'m2and it will be observed that timber falls
within the lower part of this range. However, it should be appreciated that whilst
timber in small clear specimens can be regarded as a very tough material, the
inclusion of knots as occurs in much structural timber, markedly reduces its
toughness and, when present to a considerable extent, can render the timber
brittle in behaviour.

Table 9. Toughness of materials


Work of fracture
(J m')
Ductile metals 100
GRP 105
Wood 10'
CFRP 10'

24
Timber review
Fracture morphology
With the advent of electron microscopy it has become fairly well established
that failure in tension occurs longitudinally between the S, and S, layers, at least
as far as the late wood cells are concerned. In the early wood, fracture occurs
transversely across the cell wall (Nordmann ?L Qvickstrom, 1969) and it is now
considered that these thin walled cells contribute very little to the tensile strength
of the timber. The contrast in failure mode between the longitudinal shearing
action within the cell wall of the late wood cells and the very short transverse
fracture of the early wood cells is illustrated in Fig. 9.
Since the tensile strength of individual cells in tension is considerably greater
than their aggregate strength it might be assumed that, in tensile stressing,
separation of complete cells would occur through failure of the intercellular
cement. This does not appear to be the case and microscopic observations have

Fig. 9. Tensile failure in spruce showing transverse cell failure of the


early wood (left) and longitudinal intra wall shear failure of the late
wood cells (right). ( x 110, polarized light).
25
J.M.Dinwoodie
been supported by mathematical modelling. Using orthotropic analysis Mark
(1967) has calculated the theoretical strengths of the various cell-wall layers and
has shown that the direction and level of shear stress in the various wall layers
was such as to initiate failure between the S, and S, layers. Mark's treatise has
received a certain amount of criticism on the grounds that he has treated one cell
in isolation, opening it up longitudinally in his model to treat it as a two-dimen-
sional structure; nevertheless, the work marked the beginning of a new phase of
investigation into elasticity and fracture and the approach has been modified and
subsequently developed as already described under the section on stiffness. This
extension of the work has explained the initiation of failure at the S,/S, boundary,
or within the S, layer, in terms of either buckling instability of the microfibrils
(Schniewind & Barrett, 1969) or the formation of ruptures in the matrix or
framework giving rise to a redistribution of stress.
Temperature has a marked influence on the morphology of fracture as is the
case with many other polymeric materials. Its influence is most apparent in tensile
stressing perpendicular to the grain; at normal temperature the fracture line is
most frequently found along the cell cavity though it does occur within the cell
wall, usually within the primary or S, wall layers. However, at elevated tempera-
tures fracture within the cell wall predominates (Koran, 1967).
Compression failure is a slow yielding process in which there is a progressive
development of structural change. The initial stage of this sequence appears to
occur at a stress less than 25",, of the ultimate failing load (Dinwoodie, 1968)
though Keith (1971) considers that these early stages do not develop until about
60n/,, of the ultimate. Both workers agree, however, that the sequence originates
in the tracheid or fibre wall at that point where the longitudinal cell is displaced
vertically to accommodate the horizontally running ray.
The resulting deformation assumes the form of a small kink in the microfibrillar
structure, and because of the presence of crystalline regions in the cell wall it is
possible to observe this feature using polarization microscopy. As stress and strain
increase, these kinks become more prominent and increase numerically, generally
in a preferred lateral direction, horizontally on the radial plane (Fig. 10) and at
an angle to the vertical axis of from 45 to 60 on the tangential plane. These lines
of deformation generally called a crease and comprising numerous kinks continue
to develop in width and in length and, at failure, can be observed by eye on the
face of the block of timber. At this stage there is considerable buckling of the cell
wall and delamination within it, usually between the S , and S, layers (Wardrop
& Addo-Ashong, 1964). Models have been produced to simulate buckling be-
haviour and calculated crease angles for instability agree well with observed
angles (Grossmann & Wold, 1971).
At a lower order of magnification, Dinwoodie (1974) has shown that the angle
at which the kink traverses the cell wall varies systematically between early wood
and late wood, between different species and with temperature. Almost 72",, of
the variation in shear angle could be accounted for by a combination of the angle
of the microfibrils in the S, layer and the ratio of cell wall stiffness in longitudinal
and horizontal planes.
Attempts have been made to relate the size and number of kinks to the amount
of elastic strain or the degree of plastic deformation (Keith, 1971; Dinwoodie,
1975a). Under conditions of prolonged loading, the creep function appears to
provide the most sensitive guide to the formation of cell-wall deformation (Keith,
1971, 1972).
Timber is a tough material in the absence of knots and defects and Table 9
compares the work of fracture for timber with that of other materials. Since the

26
Timber review
energy required to break the cellulose molecule is only a fraction of the total
energy to fracture it follows that the high toughness of wood must be due in part
to the high frictional forces required to break up the composite nature of the cell
wall especially under the action of shear stresses and to the crack stopping
effect of the various layers within the cell wall (Cook & Gordon, 1964) and in part
to the considerable elongation of cells following cell wall delamination and
buckling under tensile stressing (Page, El-Hosseing & Winkler, 1971; Gordon &
Jeronomidis, 1974). An example of crack stopping due to the prejence of inter-
face3 is presented in Fig. 11;the angle at which the secondary cracks have formed
coincides with the microfibrillar orientation of the S , layer of the cell wall.

Fig. 10. Formation of kinks in the cell wall of spruce timber during
longitudinal compression. ( x 990, polarized light).

Loss in toughness can arise not only on account of the presence of defects and
knots as noted earlier, but also through the effect of acid, prolonged elevated
temperatures or fungal attack on wood, or the presence of compression damage
resulting from overstressing within the living tree or in the handling or utilization
of timber after conversion (Koehler, 1933; Dinwoodie, 1971).

Other properties
There are many other physical and mechanical properties of timber which can
be related to its basic structure. The high thermal insulation of timber and the
ease with which it can be sawn and planed reflects the low density and cellular
structure of the material. The degree of impregnation of timber with artificial
preservatives or fire retardants is determined by the type and frequency of the
pitting in the cell wall, together with the physiological changes associated with
aging of the tree. Many more examples are available but are outside the scope of
this present review.
27
Table 10. Timber as a material
Advantages Disadvantagcs
Property Structural reason Property Structural reason
(1) High tensile strength Longitudinal covalent bonding in (1) Variability in performance Genetical variation between trees
unit cell combined with environmental
High intercellular shear resistance influences
(2) Very high specific tensile As above, plus low density and (2) Strength loss with increasing Chemical composition and struc-
strength and stiffness composite structure moisture content ture of cell wall
( 3 ) High toughness Composite structure containing ( 3 ) Dimensional instability in As above plus arrangement of
interfaces presence of moisture gradients cells and orientation of micro-
fibrils
(4) Good working properties Low density, cellular structure, (4) Crccp Polymeric viscoelastic material
high stiffness
(5) High thermal insulation Cellular polymeric structure ( 5 ) Anisotropv Cellular arrangement and micro-
fibrillar orientation
(6) Continued availability Renewable resource (6) Degradation by fungi and Organic material
insects
(7) Low cost relative to othcr Low energy requirements (7) Combustible (but can be Organic material
materials treated with fire retardants)
Timber review

Fig. 11. Crack-stopping in a fractured rotor blade. The orientation of


the secondary cracks corresponds to the microfibrillar orientation of the
middle layer of the cell wall. ( x 990, polarized light).

CONCLUSION
The structure of this polymeric, cellular composite has been examined at four
levels of order and a whole series of physical and mechanical properties of this
material have been related to its intriguing and complex structure; the most
important relationships are summarized in Table 10 which also attempts to assess
the general advantages and drawbacks of using this particular material. Although
competition is growing from the light-weight metals on the one hand and from
more stable and stiffer plastics on the other, the demand for timber is unlikely
to lessen due primarily to its low initial cost relative to other materials and its
unique combination of mechanical properties.

References
Anon. (1974) U.K. wood and lumber imports, 1973. Timb. TradesJ. 289, 34.
Armstrong, L.D. & Christcnsen, G.N. (1961) Influence of moisture changes on deformation
of wood under stress. Nature, Lond. 191, 869.
Bailey, I.W. & Vestal, Mary (1937) T h e orientation of cellulose in the secondary wall of
tracheary cells. 3. Arnold Arbor. 18, 185.
Bannan, M.W. (1954) Ring width, tracheid size and ray volume in stem wood of Thuja
occidentalis. Can. 3 . Bot. 32, 466.
Bannan, M.W. (1955) T h e vascular cambium and radial growth in Thuja occidentalis. Can.
3 . Bot. 33, 113.
Bannan, M.W. (1960) Cambial behaviour with reference to cell length and ring width in
Thuja occidentalis. C a n . 3 . Bot. 38, 177.
Barber, N.F. (1968) A theoretical model of shrinking wood. Holzforschung, 22 (4), 97.
Barber, N.F. & Meylan, B.A. (1964) T h e anisotropic shrinkage of wood. A theoretical model.
Holzforschung, 18, 146.
Barkas, W.W. (1946) T h e anisotropic elastic properties of wood. Confr. British Rheologist Club.
Barkas, W.W. (1949) The Smelling of Wood Under Stress. HMSO.

29
J. M. Dinzoodie
Barrett, J.D. & Schniewind, A.P. (1 973) Three-dimensional finite-element models of cylin-
drical wood fibers. Wood Fiber, 5, 215.
Barrett, J.D., Schniewind, A.P. & Taylor, R.L. (1972) Theoretical shrinkage model for wood
cell walls. Wood Sci. 4, 178.
Bolton, A.J. & Petty, ].A. (1975) Structural components influencing the permeability of
ponded and unponded Sitka spruce. 3. Microsc. 104, 33.
Carruthers, J.F.S. & Hanson, O.P. (1974) Review and outlook for wood and substitute
materials. Tenth Commonwealth Forestry Conference.
Cave, I.D. (1968) T h e anisotropic elasticity of the plant cell wall. Wood Sci. Technol. 2, 268.
Cave, I.D. (1972) A theory of the shrinkage of wood. Wood Sci. Technol. 6,284.
Chang, M. (1971) Folding chain model and annealing of cellulose. J . Polym. Sci. C36, 342.
Christensen, G.N. (1967) Sorption and swelling within wood cell walls. Nature, Lond.
213, 782.
Chow, S. (1973) Molecular rheology of coniferous wood tissues. Trans. SOC.Rheol. 17, 109.
Clouser, W.S. (1959) Creep of small wood beams undcr constant bending load. F P L Report
2150, p. 25.
Cook, J. & Gordon, J.E. (1964) A mechanism for the control of crack propagation in all
brittle systems. Proc. R. SOC.A282, 508.
C M , W.A. (1967) Wood Ultrastrucrure-An Atlas of Electron micrographs. University of
Washington Press.
Cowdrey, D.R. & Preston, R.D. (1966) Elasticity and microfibrillar angle in the wood of
Sitka spruce. Proc. R. SOC.Lond. B166,245.
Davidson, R.W. (1962) T h e influence of temperature on the creep in wood. Foresr Prod. 3.
12, 377.
Dennis, D.T. & Preston, R.D. (1961) Constitution of cellulose microfibrils. Nature, Lond.
191,667.
Dinwoodie, J.M. (1965) Tensile strength of individual compression wood fibres and its
influence on properties of paper. Nature, Lond. 205, 763.
Dinwoodie, J.M. (1968) Failure in timber. Part 1 . Microscopic changes in cell-wall structure
associated with compression failure. J. Inst. Wood Sci. 21,37.
Dinwoodie, J.M. (1971) Brashness in timber and its significance. J . I x s r . Wood Sci. 28, 3.
Dinwoodie. J.M. (1974) Failure in Wood. Part 2. T h e angle of shear through thc cell wall
during longitudinal compression stressing. Wood Sci. 7 ' e c h ~ ~8,l . 56.
Dinwoodie, J.M. (1975a) Failure in Wood. Part 3. T h e distribution and structure of slip
planes and their relationship with failure strain. ( I n preparariou.)
Dinwoodie, J.M. (1975b) Failure in Timber. Part 4. The effect of' longitudinal compression
on some mechanical properties. ( I n preparariou.)
Draffin, J.O. & Miihlenbruch, C.W. (1937) A S T M Report 96.
Dunning, C.E. (1969) T h e structure of longleaf-pine latewood. 1. Ccll-wall morphology
and the effect of alkaline extraction. TAPPI, 52, 1326.
Echenique-Manrique, R. (1967) Stress relaxation of wood at several levcls of strain. Tech. Rep.
Yale Sch. For. No. 37, pp. 29.
El-osta, M.L.M. & Wellwood, R.W. (1972a) Short-term creep as related to microfibril
angle. Wood Fiber, 4, 26.
El-osta, M.L.M. & Wellwood, R.M. (1972b) Short-term creep as related to cell-wall
crystallinity. Wood Fiber, 4, 204.
Food and Agricultural Organisation Yearbook 1969,70.
Frey-Wyssling, A. & Miihlethaler, K. (1963) Die Elcmentarfibrillen der Cellulose. Makro-
molek. Chem. 62,25.
Frey-Wyssling, A. & Miihlethaler, K. (1965) Ultrasrriictural Plant Cyrology, p. 34. Elsevier,
New York.
Frey-Wyssling, A., Miihlethaler, K. & Muggli, R. (1966) Elementary fibrils as structural
elements of native cellulose. Holz. Rho- u. Werkstofl, 24,443.
Frey-Wyssling, A. (1968) T h e ulstrastructure of wood. Wood. Sci. Technol. 2, 73.
Gibson, E.J. (1965) Creep of wood: rBle of water and effect of a changing moisture content.
Nature, Lond. 206, 213.
Gordon, J.E. & Jeronimidis, G. (1974) Work of fracture of natural cellulose. Nature, Lond.
252, 116.
Goring, D.A.I. & Timell, T.E. (1962) Molecular weight of native cellulose. T A P P I , 45, 454.
Grossman, P.U.A. & Wold, M.B. (1971) Compression fracture of wood parallel to the grain.
Wood Sci. Technol. 5, 147.
Harada, H. (1965) Ultrastructure and organization of gymnosperm cell walls. In: Cellular
Ulrrastrucrure of Woody Planrs (Ed. by W.A. CBtC.), p. 215. Syracuse University Press.

30
Timber review
Harris, J.M. & Meylan, B.A. (1965) The influence of microfibril angle on longitudinal and
tangential shrinkage in Pinus radiata. Holzforschung, 19, 144.
Hearle, J.W.S. (1963) The fine structure of fibres and crystalline polymers. 1. Fringed fibril
structure. J. appl. Polym. Sci. 7, 1175, 1193.
Hearmon, R.F.S. (1948) The elasticity of wood and plywood. Spec. Rep. Forest Prod. Res.
7,a7.
Hearmon, R.F.S. (1953) The elastic and plastic properties of natural wood. In: Mechanical
Properties of Wood atid Paper, p. 19. R. Meredith, Amsterdam.
Hearmon, R.F.S. & Paton, J.M. (1964) Moisture content changes and creep of wood. Forest
Prod. 3. 14, 357.
Hermans, P.H., De Booys, J. & Maan, C.H. (1943) Form and mobility of cellulose molecules.
Kolloid. Z. 102, 169.
Hcyn, A.N.J. (1969) T h e elementary fibril and super-molecular structure of cellulose in
softwood fiber. J . Ulrrastruct. Res. 26, 52.
Hillis, W.E. (1962) Wood Iktractivzs and Their Significance to the Pulp and Paper Industry.
New York, Academic Press.
Ifju, G. (1964) Tensile strength behaviour as a function of cellulose in wood. Forest Prod.3.
14, 366.
Jaswon, M.A., Gillis, P.P. & Marks, R.E. (1968) T h e clastic constants of crystalline native
cellulose. Proc. R . SOC.A306, 389.
Keith, C.T. (1971) The anatomy of compression failure in relation to creep-inducing stress.
Wood Sci. 4, 71.
Keith, C.T. (1972) T h e mechanical behaviour of wood in longitudinal compression. Wood
Sci. 4, 234.
Kelsey, Kathleen E. (1963) A critical review of the relationship between the shrinkage and
structure of wood. Tech. Pap. C S I R O Aust. Div. For. Prod. 28.
King, E.G. J r (1961) Time-dependent strain behaviour of wood. Foresr Pr0d.J. 11, 156.
Kingston, R.S.T. & Budgen, Beverley (1972) Some aspects of the rheological behaviour of
wood. Part IV: Non-linear behaviour at high stresses in bending and compression.
Wood Sci. Technol. 6 , 230.
Kingston, R.S.T. & Clarke, L.N. (1961) Some aspects of the rheological behaviour of wood. I
T h e effect of stress with particular reference to creep. 2. Analysis of creep data by
reaction-rate and thermodynamic methods. Aust. 3. appl. Sci. 12, 21 1.
Klauditz, W. (1952) Zur biologisch-mechanischen Wirkung des Lignins im Stammholz der
Nadel- und Laubholzer. Holzforschung, 6, 7C.
Koehler, A. (1933) Causes of brashness in wood. Bull. Forest Prod. Lab. (Madison),342, pp. 40.
Kollmann,F. (1961) Rheology and structural strength of wood. Holz. Roh- u. Werkstoff. 19,73.
Kollmann, F.F.P. & CBtk, W.A. (1968) Principles of Wood Science and Technologv. 1 . Solid
wood. Springer-Verlag, Berlin.
Kollmann, F. & Krech, H. (1960) Dynamische Messungen der elastischen Holzeigenschaftcn
und der Dampfung. Holz. Roh- u. Werkstoff, 18, 41.
Koran, Z. (1967) Electron microscopy of radial tracheid surfaces of black spruce separated
by tensile failure at various temperatures. T A P P I , 50, 60.
Lavers, Gwendoline M. (1969) T h e strength properties of timbers. Bull. Forest Prod. Res.
Lab. 50. 2nd edn.
Liang, C.Y. & 'Marchessault, R.H. (1959) Hydrogen bonds in native cellulose. 3. P o b m . Sct.
35, 529.
Licse, W. (1963) Tertiary wall and warty layer in wood ce1ls.J. Polym. Sci. CZ, 213.
McIntosh, D.C. (1965) Wall structure of loblolly pine summer wood holocellulose fibres in
relation to individual fibre strength. In: Cellular Ultrastructure of Woody Plants.
Syracuse University Press.
Manley, R. St J. (1964) Fine structure of native cellulose microfibrils. Nature, Lond. 204, 1155.
Mark, H. (1952) Cellulose : Physical evidence regarding its constitution. In Wood Chemistrjl
(Ed. by L.E. Wise and E.C. Jahn), p. 132. Reinhold, New York.
Mark, R.E. (1967) Cell Wall Mechanics of Tracheids. Yale University Press, New Haven.
Mark, R.E., Kaloni, P.N., Tang, R.C. & Gillis, P.P. (1969) Cellulose: Refutation of a folded
chain structure. Science, 164, 72.
Mark, R.E. & Gillis, P. P. (1970) New models in cell-wall mechanics. Wood Fiber, 2, 79.
Meyer, K.H. & Mark, H. (1930) Der Aufbau der Hochpolymeren Organischen Naturstoffe.
Leipzig.
Meyer, K.H. & Misch, L. (1937) Position des atomes dans le nouveau modele spatial de la
cellulose. Helv. chim. Acta, 20, 232.
Meyer, K.H. (1950) Natural and synthetic high polymers-High Polymer Ser. Vol. IV.
Interscience, New York.

31
J . M . Dinwoodie
Meylan, B.A. (1972) T h e influence of microfibril angle on the longitudinal shrinkage-
moisture content relationship. Wood Sci. Technol. 6,293.
Nieduszynski, I. & Preston, R.D. (1970) Crystallite size in natural cellulose. Nature, Lond.
225,273.
Nissan, A.H. & Sternstein, S.S. (1961) Cellulose as a viscoelastic material. Proc. Wood Chem.
Symp. ( I U P A C ) , Montreal, p. 131. Butterworths, London.
Nordrnan, L.S. & Qvickstrom, B. (1969) Variability of the mechanical properties of fibers
within a growth period. In: The Physics & Chemistry of Wood Pulp Fibres (Ed. by
D. H. Page). S.T.A.P. No. 8, 177.
Page, D.H., El-Hosseiny, P. & Winkler, K. (1971) Behaviour of single wood fibres under
axial tensile strain. Nature, Lond. 229, 252.
Parratt, N. J. (1972) Fibre Reinforced Materials Techriology. p. 2. Van Nostrand Reinhold,
London.
Pearson, R.G. (1972) T h e effect of the duration of load on the bending strength of wood.
Holzforschung, 26, 153.
Pearson, R.G. (1974) Application of fracture mechanics to the study of the tensile strength
of structural lumber. Holzforschung, 28, 11.
Preston. R.D. (1947) T h e fine structure of the wall of the conifer tracheid. 11. Optical
properties of dissected walls in Pinus insignis. I’roc. K . SOC.B134, 202.
Preston, R.D. (1959) Wall organization in plant cells. In: Int. Rev. Cyrol. VIII, 33.
Preston, R.D. (1964) Structural and mechanical aspects of plant cell walls. In: The Formation
of Wood in Foresr Trees (Ed. by H. M. Zimmcrmann), p. 169. Academic Press, New
York.
Preston, R.D. (1971) Negative straining and cellulose microfibril size. 3 ’ . Microsc. 93, 7.
Preston, R.D. & Wardrop, A.B. (1949) The fine structure of the wall of the conifer tracheid,
IV Dimensional relationships in the outer layer of the secondary wall. Biochim.
Biophys. Acta, 3, 585.
Price, A.T. (1928) Mathematical discussion on structure of wood in relation to its elastic
properties. Phil. Trans. R . SOC.Lond. A228, 1.
Roelofsen, P.A. (1959) The Plant Cell Wall. Borntragcr, Berlin.
Sadoh, T. & Kingston, R.S.T. (1967) Longitudinal shrinkage of wood. Pt 11. The relation
between longitudinal shrinkage and structure. Wood S c i . Technol. 1, 81.
Schniewind, A.P. (1968) Recent progress in the study of the rheology of wood. Wood Sci.
Technol. 2, 188.
Schniewind, A.P. & Barrett, J.D. (1969) Cell wall model with complete shear restraint. Wood
Fiber, 1, 205.
Schniewind, A.P. & Barrett, J.D. (1972) Wood as a linear orthotropic viscoelastic material.
Wood Sci. Technol. 6,43.
Schniewind, A.P. & Lyon, D.E. (1973) A fracture mechanics approach to the tensile strength
perpendicular to grain of dimension lumber. Wood Sci. Technol. 7, 45.
Senft, J.F. & Suddarth, S.K. (1971) An analysis of creep-inducing stress in Sitka spruce.
Wood Fiber, 2, 321.
Stamm, A.J. & Smith, W.E. (1969) Laminar sorption and swelling theory for wood and
cellulose. Wood Sci. Technol. 3, 301.
Sullivan, J.D. (1968) Wood cellulose protofibrils. TAI’I’I, 51, 501.
Sulzberger, P.H. (1948) T h e effect of temperature on the strength properties of wood, ply-
wood and glued joints at various moisture contents. C S I R O Div. Foresr. Prod. Project
TP10-3.
Sulzberger, P.H. (1953) T h e effect of temperature on the strength of wood. Aeron. Kes. Cons.
Comm. Rep. ACA-46, Melbourne.
Sunley, J.G. (1968) Grade stresses for structural timbers. Bull. Forest. Prod. Res. 47, HMSO.
Tang, R.C. (1972) Three-dimensional analysis for elastic behaviour of wood fiber. Wood
Fiber, 3, 210.
Tang, R.C. & Hsu, N.N. (1973) Analysis of the relationship between microstructure and
elastic properties of the cell wall. Wood Fiber, 5, 139.
Thunell, B. (1941) Uber die Elastizitat schwedischen Kiefernholzes. Holz. Roh- u. Werkstoff
1, 15.
Wardrop, A.B. (1964) T h e structure and formation of the ccll wall in xylem. In: The Forma-
tion of Wood in Foresr Trees (Ed. by M.H. Zimmermann), p. 87. Academic Press.
Wardrop, A.B. & Addo-Ashong, F.W. (1964) T h e anatomy and fine structure of wood in
relation to its mechanical failure. Proc. Tewkesbury Symp. p. 169.
Wood, L.W. (1951) Relation of strength of wood to duration of load. Forest. Prod. Lab.,
Madison, Rep. No. R1916.

32

You might also like