You are on page 1of 24

23

Wood
23.1 Introduction.......................................................................................623
23.2 Uses......................................................................................................623
23.3 General Properties.............................................................................624
Wood as a Material

23.4 Degradation Agents...........................................................................626


Nonliving Agents of Degradation: A Variety of Nonliving Agents
Include Chemicals, Sunlight, and Mechanical Damage (The Damage
Can Be Either Visual or Structural) Biological Deterioration

23.5 Wood Protection................................................................................636


Natural DurabilityPreservative TreatmentNon-Biocidal
Treatments Developing New Wood Protectants with Standardized
Testing Protocols Methods of Wood Treatment

Jeffrey J. Morrell
Oregon State University

23.6 Summary............................................................................................ 643


References...................................................................................................... 644

23.1Introduction
From the time humans emerged on the plains in Africa, wood, in one form or another, has played an
important role in everyday life. From the spears that brought down game to the wood fuel used to cook
it and the poles that provided night time shelter, wood played a variety of roles in human development
(Graham 1973). As human culture developed, wood continued to provide fuel, structural materials, and
a variety of other products including food and medicinals. Over the past century, the role of wood has
changed in many parts of the globe, but it remains one of our most important renewable structural materials. In comparison with other materials, wood is strong, light, and easily worked with simple tools. In
terms of environmental impacts, wood harvest does necessitate cutting of trees; however, replanting and
careful management creates an inexhaustible resource. In addition, wood sequesters or captures carbon
from the air, thereby reducing the potential impacts of carbon dioxide releases and reducing the pace
of global climate change. At the same time, wood is a natural material that is subject to physical and
biological degradation over time and it is this susceptibility to degradation that is its weakest attribute.

23.2Uses
Despite the risk of degradation, wood is used for houses, poles, marine piling, railway ties, composite
panels, wood/plastic composites, and paper products. It is revered for its beauty and employed because
of its utility. Wood has exceptional structural and thermal properties.

623

624

Other Natural Materials

23.3 General Properties

Downloaded by [San Jose State University] at 11:34 15 April 2016

23.3.1 Wood as a Material


Wood in the living tree serves two functions. First, it supports the foliage or canopy above, allowing it
to reach above competing plants or trees to capture sunlight for photosynthesis. At the same time, the
wood is a conduit for water and nutrients to move up and down the tree. If we look at a cross-sectional
cut from a tree, we can see the bark, which protects the living tree from injury (Figure 23.1). Beneath the
bark, we can see sapwood and heartwood (USDA 1996). Sapwood is the living part of the tree that conducts fluids. Inside that zone is the heartwood, which is the collection of nonliving cells that provides
some support for the canopy above but no longer conducts fluids. If we look closer on our cross section,
we see numerous cells with large openings called lumens. These hollow tubes or cells are both conducting elements and structural support for the tree.
There are two broad groups of trees, softwood and hardwoods. These terms are misnomers because
many softwoods are quite hard, while many hardwoods are soft. Softwoods or gymnosperms comprise
the oldest tree group, and we often call them conifers or evergreens because many retain their needles
for long periods. The hardwoods or angiosperms are considered to be more advanced and have leaves in
place of needles. In addition, the anatomy of these two groups varies.
Trees in temperate regions generally produce their conductive cells (vessels or tracheids) with
larger openings or lumens early in their growing season and cells with smaller opening later as
growth slows. The larger cells are called earlywood and the later cells are called latewood. Earlywood
and latewood formed in a given season comprises one annual growth ring. Trees in tropical or subtropical climate sometimes lack visible rings or the rings may correspond to wet/dry cycles rather
than a given year.
The wood of softwoods has relatively few cell types. The primary softwood cell is the longitudinal tracheid that provides structural support and conducts fluids between the roots and needles. Longitudinal
tracheids are long, extending to several millimeters in length in some species. Ray tracheids are oriented
from the bark to the pith of center of the living tree. They are primarily used to move fluids laterally.
Ray parenchyma are similarly oriented but primarily serve to store starches, lipids, proteins, and other
compounds that might be used to regenerate foliage or roots. In addition, some softwoods contain resin
canals. These are essential holes in the wood structure surrounded by epithelial cells that produce resins. The resins are released when the tree is wounded and serve to limit entry of fungi and insects into

FIGURE 23.1 Cross section of a Douglas fir tree showing bark, sapwood, and hardwood with easily seen annual
rings consisting of dark latewood and lighter earlywood.

625

Downloaded by [San Jose State University] at 11:34 15 April 2016

Wood

thewood. These resins are often extracted from wood for production of turpentine and other products
that were formerly called naval stores because of their use in protecting wood used in ships. All cells
inthe wood are connected with one another through openings termed pits. Pits are openings in the
wood cell that allow fluids and some smaller particles to move between cells.
Hardwoods contain more cells types. The cells that transport liquids longitudinally are called vessels.
These cells are connected with one another through sieve plates that allow liquids to be transported.
Fibers are thick-walled cells that surround the vessels. They are also oriented longitudinally and provide
structural support. Hardwoods also have ray parenchyma that store nutrients. As with softwoods, individual cells are connected with one another through pits.
If we look closer at an individual cell, we can see that there are layers. In softwoods, the middle
lamella separates individual cells and is integrated with the primary cell wall layer (Figure 23.2). The
secondary cell wall is composed of three layers, the S1, S2, and the S3. These layers differ in thickness, the
percentages of polymers present, and the orientation of these polymers (Harada and Cote 1985). All cells
are composed of three primary polymers, cellulose, hemicellulose, and lignin. Cellulose is composed of
repeating units of -D-glucose. Cellulose chains vary in length but usually contain approximately 2000
glucose units in wood. Cellulose chains tend to interact with one another to form microfibrils.
Microfibrils contain areas where the cellulose chains are highly ordered and crystalline in nature as
well as other areas that are less organized and amorphous. The structure and orientation of the cellulose microfibril gives wood its exceptional longitudinal strength. It also makes wood more resistant to
degradation than other organic materials. Hemicellulose is a heteropolymer consisting of glucose, mannose, galactose, arabinose, and a host of other sugars. Unlike cellulose, which is a homopolymer and
lacks side chains, hemicellulose is highly branched and much more susceptible to degradation. Lignin
is composed of phenyl propane units with a variety of linkages between individual units. The variety
of linkages makes lignin one of the most recalcitrant natural polymers. Lignin compounds can last for
decades in soil. Lignin, cellulose, and hemicellulose are integrated in a matrix that takes advantage of
the properties of each polymer. Cellulose provides exceptional strength, lignin provides resistance to
degradation, and hemicellulose acts as an encrusting medium that binds the mixture.

S1
S2
S3

FIGURE 23.2 Wood cells typically consist of multiple layers termed the primary and middle lamella and three
layers of secondary cell wall, termed S1, S2, and S3. The orientation of the cellulose microfibrils varies within these
layers and gives wood its unique properties.

Downloaded by [San Jose State University] at 11:34 15 April 2016

626

Other Natural Materials

The arrangement of the cells makes wood exceptionally strong in tension and compression along the
longitudinal plane but weaker in other directions. These variable properties make wood anisotropic with
different properties in the tangential, radial, and longitudinal directions. In addition, wood tends to have
differential properties depending on moisture content. As wood sorbs moisture, the water binds to the cellulose microfibrils, which expand. This expansion reduces the material properties. Many wood properties
decrease as the moisture content increases from 0 to approximately 27%30%. At this point, the hydroxyl
groups on the cellulose are largely saturated and free water will begin to accumulate in the cell lumens. This
point is termed the fiber saturation point (fsp). Wood properties are largely stable above the fsp, although
the presence of free water in the wood can create other problems as will be discussed later in this chapter.
Wood cells can also contain other materials including minerals, sugars, starches, proteins, lipids, fatty
acids, and resins that are typically found in the rays in the sapwood. These materials are termed extractives
and have no structural function. As the sapwood ages, however, the cells in the sapwood begin to die and
the stored materials can be converted to a variety of other compounds including phenolics and glycosides.
The resulting wood is termed heartwood. Some materials in the heartwood can be exceptionally toxic to
various agents of deterioration, rendering the wood resistant to degradation (Hillis 1962). Minerals, such
as silica, may also increase wood hardness and improve durability against insects or marine borers. Not
all woods have durability, and the degree of durability can vary widely within heartwood of a given tree
or within trees in a given stand. Sapwood is generally considered to be nondurable, regardless of species.
During heartwood formation, the pits connecting individual cells close and become encrusted with materials that sharply reduce fluid flow. This can have important implications for drying and impregnation.

23.4 Degradation Agents


While the structure and organization of cellulose, hemicelluloses, and lignin makes wood much more
resistant to deterioration than other biological materials, a variety of physical and biological agents can
affect wood properties. Anything that negatively affects wood properties is termed deterioration and this
damage can be caused by both living and nonliving agents (Hale and Eaton 1992; Zabel and Morrell 1992).
Often, deterioration is caused by a combination of agents, making it difficult to determine an exact cause.

23.4.1Nonliving Agents of Degradation: A Variety of Nonliving


Agents Include Chemicals, Sunlight, and Mechanical Damage
(The Damage Can Be Either Visual or Structural)
The ultraviolet (UV) component of sunlight can be especially damaging to wood surfaces (Feist 1990;
Evans et al. 1996) (Figure 23.3). The energy in the UV range is released into the wood surface and can
cause the release of free radicals that then react with various wood components. Lignin is most sensitive to UV light and begins to degrade within hours of sun exposure. The weakened lignin opens up the
other wood components to degradation. The weakened wood is then susceptible to either wind or water
erosion. In general, dark woods will tend to lighten and light wood will tend to become darker with UV
exposure. While UV damage is shallow, over time, erosion of the weaker wood and further damage to
freshly exposed wood can lead to uneven wood surfaces. In some applications, however, UV-exposed
wood is considered an attractive attribute and weathered barn siding can sell for a premium. In general,
UV damage to wood is shallow and cosmetic and, by itself, only reduces wood dimensions a few mm
over a century.
Wood can also be damaged by prolonged exposure to high temperature. This effect is often exacerbated when the wood is wet. In general, hemicellulose is most sensitive to heat followed by cellulose and
finally lignin. Wood combusts in the presence of oxygen when heated over 451F, but prolonged exposure to lower temperatures (>100C) will produce darkening and strength loss.
Although wood is generally resistant to many chemicals and wood tanks are often used to store dilute
acids or bases, a variety of chemical agents can affect wood properties. Strong acids attack cellulose and

Downloaded by [San Jose State University] at 11:34 15 April 2016

Wood

627

FIGURE 23.3 UV light can damaged the wood surface; the damage is shallow as shown by the exposed wood on
this western red cedar pole.

hemicelluloses and can sharply reduce wood properties. Strong bases attack lignin, reducing wood to
a mass of cellulose fibers. Application of strong bases to wood forms the basis for the paper industry.
Acid and base degradation are often associated with specific environments. For example, application
of high-pH stucco to the exterior of a house can produce shallow degradation of the underlying wood
surface. Chemical degradation is most common in industrial environments where wood is subjected to
prolonged or repeated exposures.
Strong salts can also affect wood and this is often evident in wood used in marine environments or
in wood used in salt storage facilities. Wood will absorb saltwater and, as this water evaporates, it leaves
the salt behind. Repeated wetting with saltwater and subsequent drying will result in cells so filled with
salt that they literally burst. The presence of such fibers on the wood surface can be disconcerting, but
the damage is very shallow.
Wood can also be damaged mechanically through abrasion or repeated heavy loading. Wood stairs
often show signs of abrasion on the treads as do wood decks in high traffic areas. Wood can also be
damaged by repeated heavy loading. The best example is the railroad tie that is periodically subjected to
millions of tons of loading from passing trains. Ties can last 4070 years but eventually must be replaced
often because of mechanical damage.

23.4.2 Biological Deterioration


23.4.2.1 Requirements for Degradation Agents
Wood can also be attacked by a variety of living agents. Some of these agents enzymatically degrade
the wood polymers, while others tunnel or excavate the wood. Nearly all of these agents need to have
four basic requirements acting concurrently to cause damage: adequate temperature, adequate oxygen,
water, and a food source. Preventing deterioration requires limiting one of these four requirements.
Most wood-degrading organisms can survive at a wide range of temperatures (Zabel and Morrell
1992). For example, they may not be active at temperatures below freezing but can survive prolonged
periods at low temperatures. As temperatures increase, organisms become more active but most organisms cease activity as temperatures reach 40C45C. All wood-degrading organisms succumb to

Downloaded by [San Jose State University] at 11:34 15 April 2016

628

Other Natural Materials

exposures at 67C after 75 min. Survival at temperatures between 50C and 67C depends on the length
of that exposure and stage of the organism. For example, some fungi produce temperature-resistant
resting stages that require the prolonged heating at elevated temperatures to succumb, while an active
insect will succumb at a much lower temperature. In general, temperature lethality occurs as enzymes
and other proteins necessary for respiration are irreversibly denatured.
The majority of organisms that degrade wood are aerobic; however, most require far less oxygen than
the 20% we find in air. Oxygen serves as the terminal electronic acceptor in the electron transport chain. It
is generally not feasible to exclude oxygen to inhibit decay. The exception is wood that is continually submerged in or sprayed with freshwater. In these cases, the water will fill the wood cell lumens to the point
where oxygen is limiting. Although anaerobic bacteria will begin to attack the sapwood, the rate of attack
is generally very slow and the wood beneath remains sound. There are numerous examples of logs being
excavated from rivers or buried in mud that are found to be sound after centuries of immersion. Similarly,
untreated foundation pilings used to support buildings often perform well because the wood is driven below
the water table and is thus protected from most degradation. Foundation piling can last centuries in this
type of exposure. As noted, however, oxygen is generally not a limiting factor in decay of wood structures.
Water is essential for all living organisms. In wood, water serves as a swelling agent, it is a reactant in
the degradation of cellulose and hemicelluloses and it is essential for cell respiration. In general, water
becomes limiting in wood when the moisture content declines below 20% (wt/wt%), although there are
exceptions to that rule. Wood decay generally begins when the wood is at the fiber saturation point, but
most decay fungi have moisture optima that are closer to 40%80% moisture content. Oxygen becomes
limiting as wood moisture contents increase beyond 100%140%, depending on the wood species, the
environment, and the agent of deterioration. On the opposite end of the moisture scale, most wooddegrading organisms can survive in dry wood and some can survive for prolonged periods. A number
of wood-degrading fungi produce long-term survival structures and are able to survive for a decade or
more in dry wood. Moisture control is the most common method for limiting biological deterioration
and most buildings are designed to limit wood moisture content to less than 20%.
A nutrient source is the final requirement and this is generally provided by the wood. Some organisms can only use the proteins and free sugars that are stored in the ray cells, others use hemicelluloses
and cellulose, and still others can utilize all wood components. There are also wood degraders that do
not use wood as a nutrient, but instead mine or tunnel through it to create galleries in which to live.
These organisms then leave their nests to forage for food elsewhere. When it is not possible to limit temperature, oxygen, or water, we protect wood by modifying the nutrient source to make it either toxic or
unusable and this approach forms the basis for wood preservation.
23.4.2.2Bacteria
Bacteria are generally not significant wood degraders, but they can become important in some environments, especially those with limited oxygen and elevated moisture levels. Bacteria are single-celled
prokaryotes that do not reproduce sexually. They are among the most common organisms on the planet
and are present in virtually every environment. Some bacteria attack the pectin substances in the pits
connecting cells, making the wood more permeable. If these bacteria-attacked logs are then stored in
water, the resultant increased water absorption can lead to those logs sinking. Other bacteria are capable
of degrading the secondary wood cell wall and their damage is characterized by either erosion of the cell
wall or the production of elongated tunnels in the S2 cell wall layer. The former bacteria are termed tunneling bacteria and can be very common in submerged wood or wood in very moist forest soils. Bacterial
damage is generally shallow and most prevalent in the sapwood, but tunneling bacteria can be preservative tolerant and may play a role in initial colonization of preservative-treated wood in soil contact.
23.4.2.3Fungi
Fungi are eukaryotic, filamentous organisms and are saprotrophs on a variety of materials, including
wood. The general fungal life cycle begins with a spore that germinates to produce fungal hyphae that

Downloaded by [San Jose State University] at 11:34 15 April 2016

Wood

629

FIGURE 23.4 Molds produce pigments spores on the wood surface and typically grow on sapwood as shown on
this Douglas fir pole.

grow through the substrate secreting enzymes that degrade the material. The degradation products
then diffuse back to the fungus where they are absorbed and metabolized. There are literally hundreds
of thousands of fungal species and many have adapted to utilize one or more wood components. Woodinhabiting fungi can be classified on the basis of morphology, DNA, or the damage they cause to wood.
For the purposes of this chapter, we will consider them on the basis of the damage they cause because
that has more direct bearing on wood use.
Fungi that inhabit wood can be broadly classified as mold, stain fungi, or decay fungi. Decay fungi
can be further divided into brown, white, or soft rot fungi.
Mold fungi are classified as those fungi that produce pigmented spores on the wood surface
(Figure23.4). These fungi generally do not attack any of the wood polymers, but instead live on the
materials stored in the ray cells. The primary impact of molds on wood is visual, and those spores
can generally be brushed away. The hyphae of mold fungi are clear or hyaline and grow through the
sapwood through pits. Brushing away spores will not eliminate hyphae inside the wood. The damaged
pits increase the permeability of the wood that can affect drying rates as well as the ability to evenly
finish the surface with paints or coatings. The primary issue associated with mold, however, is public
concern about potential health risks associated with some of these fungi. Some species of molds can
produce mycotoxins, and there is concern that the presence of these fungi on wood in houses may
cause health issues. There has been tremendous speculation about the risk of molds; however, most
studies have concluded that molds are primarily allergens that can aggravate existing conditions such
as asthma. Some people are exceptionally sensitive to molds, and there is no way to predict who among
the general population will have issues with these fungi. The best approach with molds is to control
moisture levels in the wood to reduce the risk of fungal attack. Where fungal attack has occurred,
cleaning the surface with soapy water and then making sure the wood is allowed to dry below 20%
moisture content can eliminate future risk.
Stain fungi are similar to molds in that they primarily attack stored compounds in the ray cells, but
they can also begin to attack hemicelluloses and cellulose if allowed to continue to grow (Scheffer and
Lindgren 1940). Stain fungi primarily attack the sapwood, but unlike molds, the hyphae of these fungi
become covered with a brownish pigment called melanin. The brownish color inside the wood is seen
as blue on the wood surface and damage caused by these fungi is commonly called blue stain. Stain
fungi may also produce spores or masses of hyphae on the wood surface and are often found in the
same locations as molds. The primary difference between mold and stain damage is that stain damage

Downloaded by [San Jose State University] at 11:34 15 April 2016

630

(a)

Other Natural Materials

(b)

FIGURE 23.5 White (a) and brown (b) rot cause severe structural damage and are usually found on the interior
of wood products.

is not easily removed, while mold damage can be brushed away. Stain fungi alter wood appearance and
increase permeability.
Decay fungi have a more profound effect on wood properties because they depolymerize and modify lignin, cellulose, or hemicelluloses. These fungi are the most prominent cause of deterioration in
wooden structures (Duncan and Lombard 1965). Brown rot fungi preferentially attack hemicelluloses
and cellulose and leave behind a modified lignin matrix, causing up to 60%70% mass loss (Figure 23.5).
Brown rot fungi tend to attack softwoods, although they will attack all types of wood. Wood damaged
by brown rot fungi tends to be darkened with numerous breaks across the grain. Brown rot fungi tend to
be more important in structural applications because they degrade cellulose far faster than they utilize
the breakdown products. Thus, at very low mass losses, the wood will experience substantial losses in
material properties. This means that wood with decay that is not yet visible to the naked eye may have
lost 60%70% of its flexural properties.
White rot fungi utilize all three wood polymers and cause up to 95%97% weight loss. White rot
fungi leave the wood a bleached white color at the advanced stages of attack. While white rot fungi can
attack all woods, they tend to be more aggressive on hardwoods. Some white rot fungi preferentially
attack lignin and there have been efforts to use these fungi for biopulping for paper products. These
fungi have also been explored for degradation of a variety of pesticides. White rot fungi tend to cause
losses in structural properties that are proportional to the mass loss. As a result, decay is often visible
when properties decline to the point where replacement might be advisable. White rot fungi tend to be
more of a problem in tropical countries because of the prevalence of hardwood use in those areas.
Soft rot fungi tend to attack the cellulose and hemicelluloses but differ from both white and brown
rot in their mode of attack. White and brown rot fungi tend to be found away from the surface and
inside the wood where the moisture levels are more stable. Soft rot fungi tend to be found in more
extreme environments, either on the wood surface where moisture levels fluctuate or in very wet environments where low oxygen levels limit white or brown rot attack. They also tend to be more damaging to the sapwood and on hardwoods. In addition, brown and white rot fungi cause their damage
from the cell lumens outward, while soft rot fungi can either grow directly through the secondary cell
wall or erode the cell wall surface from the lumen outward (Figure 23.6). Soft rot attack causes sharp
drops in material properties and this damage becomes more important because it tends to occur on
the surface. This becomes more important for wood used in applications where bending strength is
important because most of the bending strength of the wood is found in the outer 2550 mm. Soft rot
fungi sharply reduce strength in that area, thereby reducing the effective circumference and magnifying their potential impacts.

Downloaded by [San Jose State University] at 11:34 15 April 2016

Wood

631

FIGURE 23.6 Soft rot damage is usually characterized by softening of the wood surface as illustrated by this
screwdriver penetrating into a pole.

23.4.2.4Insects
A variety of insects have evolved to use wood as either a food source or as a habitat. Members of the
orders Isoptera, Coleoptera, and Hymenoptera are among the most important agents of damage for
wood in service. Depending on the insect, damage can occur in the living tree, as the tree is processed
and dries or once it is placed in service. Understanding the biology of the insect is critical for developing
strategies that render the wood unsuitable for insect utilization.
23.4.2.4.1Isoptera
The termites or Isoptera are social insects that use wood as a food source. They are mostly confined to
those equatorial and temperate zone areas between 50 N and S latitude. Termites have a highly structured caste system that includes the queen, workers, and soldiers. It is estimated that termites cause up
to $9 billion dollars in damage to wood per year in the United States. This figure represents not only
the wood loss but the cost of repairing the damage. There are three groups of wood-damaging termites.
The Rhinotermitidae or subterranean termites live in large colonies (one to several million workers)
in soil contact. They excavate tunnels through the soil searching for wood (Figure 23.7). They can also
move upward over non-wood surfaces by constructing earthen tubes. This allows them to move over

FIGURE 23.7 Termite workers on severely attacked wood.

632

Other Natural Materials

masonry foundations to attack wood in houses. There are a variety of methods for limiting subterranean
termite attack including soil poisoning, physical soil barriers, and the use of preservative-treated wood.
Dampwood termites, as the name implies, live in very wet wood and do not require soil contact to initiate a colony. The relative size of workers tends to be much larger than those for subterranean termites,
but the colonies are much smaller, numbering only hundreds to a few thousand workers. Dampwood
termite attack is usually prevented by controlling moisture. Drywood termites live in very dry wood,
usually with less than 12% moisture content. Like most termites, they are rarely seen outside of the nest,
but their damage can be detected by the presence of their frass outside the nest. Periodically, workers
will tunnel to the surface and kick out frass that has built up inside the tunnels and these piles of frass
are sometimes the only indicator of attack. Drywood termites tend to be more of a problem in drier
climates.

Downloaded by [San Jose State University] at 11:34 15 April 2016

23.4.2.4.2Hymenoptera
The order Hymenoptera includes bees, wasps, and ants. Of the many species in this order, the most
important from a wood damage perspective are carpenter bees and carpenter ants.
Carpenter bees attack dry wood in structures and do not use the wood as a food source. The adult
female tunnels into the wood and places eggs along with food for the developing larvae. She then seals
the hole and allows the eggs to develop. Once the egg proceeds through larval stages, it pupates and
emerges through the original hole. Carpenter bees can reuse the same chambers, extending them each
season. Attack can be limited by painting the wood. Preservatives have little effect because the insects
do not consume the wood.
Carpenter ants are social insects that also do not use wood as a food source (Figure 23.8) (Hansen and
Klotz 2005). The colony starts when male and female reproductives leave a mature colony and mate. The
females then seek out a sheltered area where she lays eggs and uses her fat reserves to feed the developing
larvae. These workers then feed and raise the next batch of eggs. It is only then that the workers begin
to move outside the colony to feed on sugars, proteins, and other materials. Colonies can vary in size
from a few hundred to upward of 50,000 workers. Carpenter ants are common in temperate forests and
their tunneling plays an important role in wood recycling. They are also prevalent in houses in forested
areas as well as timbers and utility poles. Controlling carpenter ants can be difficult because they do not
consume the wood and because the entire colony can move to a new location when disturbed and that
new location may be in the same structure. Carpenter ant treatments mostly consist of barrier sprays
around structures to keep the workers from foraging inside.

FIGURE 23.8 Carpenter ant attack on a creosote-treated southern pine pole.

Downloaded by [San Jose State University] at 11:34 15 April 2016

Wood

633

The Coleoptera or beetles contained the largest number of species in the order of insects. Beetles can
attack living trees, freshly fallen trees, or even dried, finished products, depending on their moisture
requirements. While a range of beetles can attack wood products, the most important are bark beetles
(Scolytidae), ambrosia beetles, metallic wood borers (Buprestidae), long-horned borers (Cerambycidae),
and powderpost beetles (Lyctidae, Bostrichidae, and Anobiidae).
Bark beetles attack live or freshly fallen trees, laying their eggs in galleries between the bark and
wood. The larvae tunnel in this same area, cutting off the flow of nutrients to the living tree. Bark beetles
do not cause appreciable wood damage, but they are associated with blue stain fungi that can discolor
the wood, making the resulting lumber less valuable. Ambrosia beetles also attack freshly cut trees as
well as wet lumber. The females tunnel into the wood for varying depths and deposit their eggs along
with a bit of a fungus. The fungus grows into the wood and when the larvae hatch, they feed on the
fungus. The fungus stains the wood, reducing the value of the finished product. Ambrosia beetles and
bark beetles are both eliminated by drying the wood after which they do not continue to cause damage
in the finished product.
Long-horned and metallic wood borers both attack freshly fallen trees and lay their eggs on or in the
bark. The larvae tunnel inward and then mine between the bark and the wood for a period of 1 to several
weeks. They then tunnel into the wood and have life cycles that last for 120 or more years, depending
on the wood condition. These larvae can survive through the sawing process and then continue their life
cycle in the structure. They will then pupate and emerge into the building. These beetles do not reinfest
dry wood, so their damage is generally more of a nuisance than a structural issue unless there are too
many beetle emergence holes on the same board. In that case, replacing or reinforcing the board may
be necessary.
Powderpost beetles attack drier wood and are often a problem in houses, barns, and bridges. There
are three broad groups of powderpost beetles with differing requirements for wood species and moisture
content. These beetles prefer sapwood and do not penetrate materials with coatings such as paint or
varnish on the surface. The adult beetles lay their eggs on the wood surface, and the larvae then tunnel
inside. The initial entry hole can be very small and difficult to detect. The damage is often only discovered when the adult insect emerges from the wood through a new and larger hole (Figure 23.9). In some
cases, the adults can mate and lay eggs inside the same piece of wood, and the damage can continue
until the entire sample is reduced to a powdery mass surrounded by a thin veneer of sound wood. Once
inside the wood, heat treatment (130F for 30 min) or fumigation are the only effective ways to eliminate the infestation. Neither of these processes, however, can prevent reinfestation. The best way to keep
thebeetles out without painting is to brush on a 1%2% solution of boric acid.

FIGURE 23.9 Powderpost beetles produce a powdery frass on the interior of the wood while leaving a thin veneer
of solid wood on the surface.

Downloaded by [San Jose State University] at 11:34 15 April 2016

634

Other Natural Materials

FIGURE 23.10 Limnoria or gribbles are free-swimming crustaceans that attack wood surfaces in salt or brackish water.

23.4.2.5 Marine Borers


Wood submerged in marine environments (i.e., salt or brackish water) is usually too wet for substantial
fungal or insect attack, but it can be damaged by a number of marine wood borers. Limnoria, shipworms, and pholads are the three most important marine wood borers. In the age of wooden sailing
vessels, marine borers were a major concern for seafarers and many ships that sank did so because of
marine borer attack. Marine borers are less of a concern these days, but they still cause substantial damage to coastal infrastructure.
Limnoria or gribbles are small (36 mm long), mobile crustaceans that tunnel in wood (Figure 23.10).
While the tunnels they produce are small, they weaken the wood surface to the point where wave action
can wear away the surface. Over time, Limnoria attack can result in substantial wood loss and collapse
of a piling.
Shipworms are mollusks that begin life as free-swimming larvae. Eventually, the larvae settle on a
wood surface and burrow inward. As they burrow, they become wormlike with a pair of rasping shells
at the head that tunnel into the wood and a pair of siphons on the opposite end that extend outward
from wood surface to exchange oxygen and nutrients with the water. The shipworm withdraws into the
hole at any sign of danger, leaving little sign of its presence. Inside, however, the shipworm removes large
quantities of wood and can reach 25 mm in diameter and 0.31.5 m in length (Figure 23.11). Eventually,
so much wood is removed that the structure collapses.
Pholads are also mollusks that begin life as free-swimming larvae. However, pholads do not use wood
as a nutrient source and they remain within their shells near the wood surface. Like Limnoria, pholads
weaken the wood surface, leading to gradual loss of cross section and eventual failure. Pholads tend to
be more prevalent in subtropical to tropical waters.
Marine borer attack is generally prevented by the use of naturally durable or preservative-treated
woods. Barriers such as polyurethanes can also protect wood but must remain intact in order to remain
effective.
23.4.2.6 Birds and Mammals
Although they do not use wood as a food source, some birds and mammals can cause wood damage.
Woodpeckers use wood poles and timbers as nesting sites. Since they have poor senses of taste and smell,
these birds are not dissuaded by the presence of preservative treatments and can cause extensive damage
to utility poles. The resulting cavities expose untreated wood to water, fungal spores, and insect attack.

Downloaded by [San Jose State University] at 11:34 15 April 2016

Wood

635

FIGURE 23.11 Shipworms are wormlike mollusks that attack the interior of wood exposed in salt or brackish water.

The result can be a small hole leading to extensive decay pockets (Figure 23.12). Many utilities wrap their
poles in metal hardware cloth to discourage attack.
A variety of animals ranging from voles to horses can chew treated wood in search of salts. As with
woodpeckers, these animals are not discouraged by treatment. Instead, farmers use untreated wood
where chewing is likely to occur or they use oil-based treatments that are less susceptible to attack.
While bird or mammal attack can be costly in individual instances, it is generally not widespread.

FIGURE 23.12 Woodpecker galleries can allow water, fungi, and insects to enter untreated wood.

636

Other Natural Materials

23.5 Wood Protection

Downloaded by [San Jose State University] at 11:34 15 April 2016

The best approach to keeping wood sound is to limit one of those four previously listed essential requirements where the most common approach is to limit moisture. Engineers and architects avoid wood
wetting by creating drainage planes that shift water away from a structure. Drainage planes can be
steep sloping roofs with long overhangs that limit the risk of wetting of the siding. They include the use
of caulking and coatings to limit moisture entry between joints, gutters to channel water away from
the structure, and foundations to lift wood out of soil contact. It is important to note that wood, even
wood that is integrated in other materials, remains susceptible to degradation. For example, the wood
in a woodplastic composite will decay under the proper moisture conditions (Mankowski and Morrell
2000). Where moisture control is not possible, the wood must be either naturally durable or be supplementally protected by chemicals.

23.5.1 Natural Durability


As mentioned earlier, the heartwood of some species contains a potential array of chemicals that render
the wood resistant to various agents of deterioration (Hillis 1962; Scheffer and Cowling 1966; Taylor
etal. 2002). The use of naturally durable woods is attractive from a public perception perspective, but
it is important to remember that the compounds that provide this protection are still potent antimicrobials. It is also important to note that natural durability varies by position in the tree and between
individual trees. The most durable heartwood is found in the base of the tree at the heartwoodsapwood
interface. Durability declines inward toward the pith and upward in a given growth ring. There is also
evidence that natural durability declines in second-growth trees of some species, notably teak and redwood. Naturally durable wood from temperate climate species tends to perform best in non-soil contact
applications. Many tropical species also perform well in soil contact, but care should be taken to ensure
that these materials originate from sustainably managed forests.
While naturally durable woods remain attractive for many applications, the supply does not meet
the demand for durable wood products. Supplemental treatment, usually by pressure impregnation,
provides an alternative method for prolonging the useful life of wood products used under adverse
conditions.

23.5.2 Preservative Treatment


The goal of wood treatment is to deliver a sufficient quantity of protectant to a depth sufficient to protect
the product over its useful life (Mac Lean 1952; Hunt and Garratt 1967). The amount of chemical deposited is usually expressed on either a weight per unit mass (pounds per cubic foot or kilograms per cubic
meter) or on a mass/mass basis, depending on the country. The depth to which the chemical is delivered
is generally expressed as either an absolute depth or a percentage of the sapwood.
Wood protectants can be either biocides that inhibit or are toxic to the various agents of decay or
they may merely alter the wood to change moisture-holding capabilities or make the wood unrecognizable to those organisms. In most countries, biocides are regulated by their governments. For example,
the U.S. Environmental Protection Agency (EPA) regulates pesticides under authority of the Federal
Insecticide, Fungicide, and Rodenticide Act (FIFRA). The EPA reviews all pertinent health and safety
data to make sure that the risks of use do not outweigh the benefits. The best example of this process
was the 1970s review of all wood preservatives (USDA 1980), but the EPA continually reappraises all
registered pesticides.
Preservatives can be formulated as either oil- or water-soluble systems, depending on the chemical
characteristics of the biocide as well as the desires of the treater (Freeman et al. 2003; Schultz et al. 2007).
Oil-based systems are preferred in some applications because they tend to make the wood more water
repellent and this can reduce the effects of weathering. They are also used on utility poles because the

Downloaded by [San Jose State University] at 11:34 15 April 2016

Wood

637

oil makes the pole easier to climb. However, oil-based systems are more costly and they make it difficult
to apply additional layers of paints or other coatings. This makes them less attractive for applications
such as decks and other residential applications. Waterbornes are widely used for residential applications because the resulting treated wood can be painted or coated and lacks unsightly surface deposits.
Currently used oil-based preservatives include creosote, pentachlorophenol, and copper naphthenate.
Creosote is the oldest wood preservative currently in use and was first patented in 1836 (Graham 1973).
It was originally produced as a by-product of coking coal for steel. Gases released as the coal was heated
were captured and condensed to produce various tars and creosote. Creosote is a mixture of polycyclic
aromatic hydrocarbons and its composition will vary depending on source. For this reason, the preservative is largely defined by the percentages of its compounds with various boiling point ranges. Creosote
is effective against fungi, insects, and most marine borers. Creosote remains widely used for railroad
ties but is also used to treat marine piling and utility poles. It is classified as a restricted use pesticide
in North America and can only be applied by those who are licensed by the appropriate state or federal
agency. Use of wood treated with creosote is not restricted, although some regulatory bodies restrict its
use over or in aquatic environments.
Pentachlorophenol (penta) was developed in the 1930s as the first synthetic organic wood preservative. It is widely effective against fungi and insects. Penta was an attractive alternative to creosote
because it was cheap and available in steady supply, while creosote supplies tended to vary with the
health of the steel industry. The primary negative issue with penta was the presence of the contaminant
dioxin that was produced as a by-product of the chlorination process. A series of reviews by the U.S.
EPA has led to penta being listed as a restricted use pesticide. Use of this chemical has been banned or
severely restricted in many countries; however, continuing data development in the United States has
shown that the chemical can be used safely when the manufacturing process was modified to reduce
the dioxin content. The primary uses for penta in the United States include utility poles, bridge timbers,
and foundation piling.
Copper naphthenate (CuNap) was developed at the end of the nineteenth century but saw little use as
a wood preservative until the 1980s. This preservative is a complex between copper and naphthenic acid
that is derived from petroleum. CuNap has a lower toxicity profile than penta, and numerous field trials
have shown that it is just slightly less effective than penta in soil contact. The primary use for CuNap is
treatment of utility poles although it can be used wherever penta is employed.
In addition to these heavy-duty wood preservatives, a variety of organic preservatives can be dissolved
in either oil or water, depending on how they are formulated. These include tributyltin oxide (TBTO),
3-iodo-2-propynyl-butylcarbamate (IPBC), tebuconazole, propiconazole, and didecyldimethylammonium chloride (DDAC).
TBTO and IPBC are effective against most fungi and insects when used away from direct soil contact.
Both are used in their oil-soluble forms for treatment of window frames. Tebuconazole and propiconazole are triazole compounds that have exceptionally low toxicity to nontarget organisms. These systems
can be formulated in oil for application to window frames, while DDAC is used as a co-biocide with
IPBC in some window treatments. The lower toxicity of these components makes them less effective in
direct soil contact.
Water-based preservatives: The most common heavy-duty water-based preservatives contain one or
more heavy metals. Copper is common to most of the systems that include chromated copper arsenate
(CCA), ammoniacal copper zinc arsenate (ACZA), ammoniacal copper quaternary (ACQ), and copper
azole (CA) (Freeman and McIntyre 2008). All of these systems use copper as the primary biocide but
incorporate a supplemental biocide to protect against copper-tolerant organisms. CCA was developed in
India in the 1930s and became the most widely used preservative in the world. The chromium in CCA
reacts with the wood as well as the copper and arsenic to help immobilize the preservative (Dahlgren
and Hartford 1972a,b,c; Dahlgren 1974). CCA is currently used to treat utility poles, marine piling,
timbers, and a range of other products for nonresidential applications. It is listed as a restricted use
pesticide.

Downloaded by [San Jose State University] at 11:34 15 April 2016

638

Other Natural Materials

ACQ and CA were developed in the late 1980s and early 1990s. Both use copper solubilized in either
ammonia or ethanol amine coupled with either DDAC or tebuconazole as a co-biocide. The copper is
immobilized as the ammonia or amine evolves from the wood. In addition to these solubilized formulations, there have been recent developments to create stable particulate systems of both ACQ and CA.
These systems are created by grinding the solid material into fine particles that remain suspended in
the treatment solution. One advantage of this approach is that the copper is less likely to leach; however,
there is a fierce debate about the effects of using non-soluble copper on preservative performance. ACZA
was developed in the 1930s specifically for the treatment of refractory or difficult to treat woods and
uses ammonia to solubilize the copper, zinc, and arsenic. The ammonia also helps to swell the wood and
remove debris from the pits, improving treatment. ACZA is primarily used in the Western United States
for treatment of Douglas fir for utility poles, marine piling, and timbers.
In addition to heavy metal-based systems, boron and fluoride have long been used for wood protection. Both chemicals are water soluble and one of their primary advantages is their ability to diffuse
into wood that is above the fsp. Conversely, however, both chemicals are susceptible to leaching if they
are used in wetter environments. Fluoride is primarily used as a supplemental treatment that is applied
to wood that is already in service; however, boron is also used as an initial treatment for wood that is
exposed inside buildings where it is unlikely to be subjected to wetting. The primary goal in these cases
is to provide protection against insects, particularly termites.

23.5.3 Non-Biocidal Treatments


Concerns about the safety hazards of pesticides have sparked increasing interest in other methods of
wood protection. Most of these strategies involve altering the wood to change its moisture-holding
characteristics or make it unrecognizable to fungal enzymes. The two most important approaches are
acetylation and heat treatment. Acetylation involves treating the wood with agents that react with the
hydroxyls present on cellulose and hemicelluloses, thereby making the wood less able to absorb water.
Acetylation is currently used in Europe and has been explored in North America. The primary limit has
been cost since effective acetylation requires substantial amounts of treatment to be effective (20%30%
weight gains). Heat treatment involves heating wood for long periods in a limited oxygen atmosphere.
The primary effect of heating is to reduce the amount of hemicelluloses and free sugar in the wood and
this in turn limits the ability of many fungi to colonize the wood. The heating can also reduce the material properties of the final product. Heat treating is used in parts of Europe, but it has not performed as
well in North America.

23.5.4Developing New Wood Protectants with


Standardized Testing Protocols
Developing a new wood protectant requires extensive testing to ensure that the system will perform for
prolonged periods of time. The U.S. EPA does not assess the efficacy of a preservative. This role is instead
supported by the American Wood Protection Association (AWPA) (2009), a nongovernmental body
whose members consist of chemical producers and users as well as those with a general interest in the
subject. The AWPA has developed a series of protocols for standardization that lists the various tests for
consideration by the technical committees. The protocol is voluntary, but it provides guidance for those
contemplating standardizing a system. Among the standards that are especially important for this process is the E10 soil block test, which is the first step in assessing a new chemical under controlled laboratory conditions (Figure 23.13). Next, a proponent would treat small stakes with varying amounts of
applied chemicals and then expose these in soil contact under AWPA Standard E7. At the same time, the
proponent might treat and expose other samples to various aboveground exposures under Standards
E15or E22 with the goal of developing data for specific applications such as decking or window frames
(Figure 23.14). The proponent would also need to develop data on the ability of the system to migrate out

Downloaded by [San Jose State University] at 11:34 15 April 2016

Wood

639

FIGURE 23.13 Example of a soil block test that is the one of the first standard methods used in North America to
assess the efficacy of a new wood preservative.

FIGURE 23.14 Example of a l-joint test that is used to assess the efficacy of preservatives for aboveground wood
protection.

640

Other Natural Materials

from the wood, cause corrosion of metal fasteners, and/or to impact wood properties. Once a system has
been proven to be effective, the proponent must develop methods for analyzing the active ingredients
in the wood and must develop data showing that wood of various species can be treated to the target
retention levels under commercial conditions.
The process of developing a wood protectant can take 35 years. Some proponents shorten this time
by going directly to the building codes without AWPA Standardization through the International Code
Council Evaluation Service (ICC-ES). ICC-ES reports can be accepted by building officials, but the level
of scrutiny to which these systems have been subjected is weaker and the processes by which treatment
quality is verified are not public.

Downloaded by [San Jose State University] at 11:34 15 April 2016

23.5.5 Methods of Wood Treatment


By its nature, wood is a differentially permeable medium, and flow through this material is primarily driven by the smallest openings that are pits. Treatment tends to be easiest longitudinally and is
often several orders of magnitude lower in the radial and tangential directions. Sapwood is more easily
treated, reflecting a tendency for the pits in the heartwood to be either closed or occluded with extractives that limit fluid flow. The methods used to deliver chemical into wood depend on the wood species,
the chemical involved, the deterioration hazard to which the wood will be exposed, and the length of
time the product is expected to perform.
The two primary methods for wood treatment are pressure and non-pressure processes (Mac Lean
1952). Non-pressure processes included brushing, spraying, dipping, and soaking. These processes are
primarily designed for surface protection and the depth of chemical penetration is generally shallow. As
might be expected, the best treatment will occur along the grain, while treatment depth in the radial or
tangential directions may extend to no more than a millimeter or two.
Dipping is used extensively in North America for treatment of wood used in windows and door frames.
Since decay in windows tends to occur at end joints, the tendency for treatment to be better in this direction and the fact that the windows are painted or clad help the relatively shallow treatment perform.
Dipping and spraying are also used for applying fungicides to freshly sawn lumber to prevent fungal stain
and mold. In this case, the treatment provides a shallow, prophylactic barrier against attack by fungal
spores that is only needed for 16 months when the wood has dried and is no longer at risk of fungal attack.
The vast majority of wood treated in North America is subjected to some combination of vacuum
and/or pressure treatment designed to drive chemicals more deeply into the wood. These processes are
usually performed in a vessel called a retort. Attached to the retort are pressure and vacuum pumps
along with various storage tanks from which chemicals can be transferred (Figure 23.15).
There are three basic treatment processes: vacuum, full cell, and empty-cell processes (Figure 23.16).
In the vacuum process, the wood is placed into the retort, the door is closed, and a vacuum is drawn
over the wood. The length of time the vacuum is drawn depends on the wood species and the treatment.
The chemical is then added into the retort, the vacuum is released, and the treatment solution is then
drawn into the wood. This process may be repeated one additional time for a vacvac process. This process produces acceptable treatment of dry sapwood and is used in Europe for treating wood for use in
windows and doorframes.
The full cell process was developed in the 1830s by a British engineer, John Bethel (Graham 1973). Wood
is placed into the retort, a vacuum is drawn over the wood, and then the treatment solution is added. The
vacuum is released and then the pressure is increased to the desired level (usually 100200psi). The pressure is held until gauges show that the wood has absorbed a sufficient amount of chemical then the
pressure is released. A series of vacuums may then be drawn to relieve any residual pressure and recover
excess chemical. The full cell process is designed to produce the maximum amount of treatment solution
uptake for a given depth of treatment. It is most often used for treatment of wood for use in marine environments where high chemical loadings are required or for water-based solutions where there is little or
no solvent cost and solution strength can be adjusted to achieve a given target.

641

FIGURE 23.15 A retort or treating cylinder used to pressure impregnate wood with preservatives.

1056

Pressure
(kPa)

Pressure
(kPa)

1056

0
Vacuum
(mm Hg)

0
Vacuum
(mm Hg)

670
(a)

4
5
Time (h)

670

(b)

4
5
Time (h)

Pressure
(kPa)

1056

0
Vacuum
(mm Hg)

Downloaded by [San Jose State University] at 11:34 15 April 2016

Wood

670
(c)

Time (h)

FIGURE 23.16 Examples of (a) full, (b) Lowry, and (c) Rueping cycles used to impregnate wood with preservatives.

Downloaded by [San Jose State University] at 11:34 15 April 2016

642

Other Natural Materials

The empty-cell processes were designed for use with more costly oil-based chemicals in terrestrial
exposures. The Lowry and Rueping processes both eliminate the vacuum and use only pressure to drive
chemical into the wood. In the Lowry process, treatment solution is added; the pressure is raised and
held until the desired amount of chemical has been absorbed. In the Rueping process, pressure is raised
to 3050 psi before the treatment chemical is added and then the pressure is raised to the desired level.
In both cases, air is trapped into the wood at the start of the process. This air is compressed as the pressure is raised. At the end of the process, this cylinder of compressed air expands outward, carrying with
it excess chemical. Vacuums are then used to hasten this release and recover excess solution. The emptycell processes produce much lower retentions or loadings than a comparable full cell process.
The AWPA provides the primary standard for pressure treatment in the United States. These standards
are based upon the hazard to which the product will be exposed (called the Use Category) and the commodity to be treated. The AWPA Standards provide minimum target retentions and specify the depth to
which the treatment must be delivered for a given species and decay risk. There are five Use Categories
beginning with the lowest risk of wood inside a house protected from wetting (Use Category1) to Use
Category 5 where the wood is exposed in a tropical marine environment. The Canadian Standards differ
slightly but follow a similar approach for hazard classes. Assessing the risk of deterioration can be complicated, but there are a number of guides for the process based upon prior performance data or climatic
conditions (Scheffer 1971, AWPA 2009).
While the treatment processes are relatively well-defined, there have been efforts to develop new methods for impregnating wood-based materials. Two recent examples are vapor phase boron and supercritical fluid impregnation. Vapor phase boron was simultaneously developed in the United Kingdom and
New Zealand (Murphy and Turner 1989) and was briefly used commercially for treatment of composite
panels. Trimethyl borate was introduced, under vacuum, into a chamber containing the wood and then
volatilized to penetrate into the panels where it reacted with water, leaving boric acid. The process was
effective but too costly to be competitive. Supercritical carbon dioxide is an excellent, non-swelling
solvent that has been shown to be able to carry a variety of biocides into wood-based materials (SahleDemessie et al. 1995). The process is used commercially in Denmark for treating window frames, but it
remains too costly for general use. These processes not only illustrate the potential for developing new
treatment methods but also demonstrate the difficulty of bringing them to widespread commercial use.
23.5.5.1 Treatment Quality
Treatment quality can be assessed under a variety of standards. For pressure-treated wood, the two
primary standards used in North America are those of the AWPA (2009) and the Canadian Standards
Association. The AWPA is one of the oldest standards-writing bodies in the world and it produces
result-oriented standards for wood treatment. Results oriented means that within certain parameters,
the treater can use any process they wish to use but they must achieve a desired outcome in terms of
chemical penetration and retention. The Canadian Standards are similarly structured. Treatment quality under these standards is usually assessed by removing increment cores from selected wood samples
in a given treatment charge, measuring preservative penetration, one for each core, and then cutting
a specific assay zone from each core. These assay zone core segments are then combined from a given
charge, ground to a dust, and quantitatively analyzed for the treatment chemical using methods that are
listed in the standards. Lumber treatments under these standards are also overseen by the American
Lumber Standards Committee (ALSC) that operates under the auspices of the U.S. Department of
Commerce. The ALSC also oversees the lumber grade stamp program. The primary goal of the quality
program is to protect the consumer.
23.5.5.2 Environmental Impacts of Treatment
While wood treatments have the positive environmental benefit of prolonging the useful life of an
already renewable material, there is also the potential for negative impacts as a result of the chemicals
employed. Most of this risk occurs at the treatment facilities where large quantities of preservatives

Downloaded by [San Jose State University] at 11:34 15 April 2016

Wood

643

are used and this is where most regulatory authorities have instituted the tightest controls. Once the
treated wood leaves the plant, there is still a risk of environmental contamination. Virtually all wood
protectants have some degree of water solubility and will migrate from the treated product into the surrounding environment. This water solubility is essential for the chemical to move into a target organism
(fungus, insect, or marine borer) to exert an inhibitory effect (Chou et al. 1971). However, problems can
arise when too much of the chemical migrates from the wood.
Migration of oil-based preservatives is a function of the degree of water solubility of the active ingredient. In most cases, the degree of migration into the soil is limited to 150300 mm away from the wood.
This reflects a combination of the relatively low levels of chemical migrating from the wood and the
tendency for the soil microflora to degrade organic compounds. As a result, treated wood poses little
risk in soil contact. The potential impact of migration from wood exposed to aquatic environments
is a function of the amount of treated wood, the water conditions, the current, and the organisms of
concern (Lebow et al. 2000). The greatest impact of treated wood in an aquatic environment occurs
shortly after installation when any excess chemical on the wood surface can be more easily dislodged.
Chemical losses then decline to a low steady state. A number of models have been developed to predict
the risk of using treated wood in aquatic environments. In general, these models show that treated wood
can pose risk when used in poorly circulating waters with minimal current but pose little risk in other
applications.
Migration of metal-based preservatives is also a function of water solubility, but the difference is
that most water-based systems are metallic and are, therefore, not degraded. Soil sampling around utility poles and fence posts, however, reveals that metal levels are elevated only for zones 150300 mm
away from the structure. In aquatic applications, metal losses are greatest shortly after installation and
decline sharply with time.
23.5.5.3 Mitigation of Environmental Risks
In response to the obvious migration of preservatives from treated wood used in aquatic environments,
treaters in some parts of the United States have developed a series of Best Management Practices (BMPs)
designed to reduce the impact of using treated wood (WWPI 2006). BMPs were primarily developed
for wood used in aquatic environments, but they also have application for wood used in other sensitive
areas. BMPs vary with the treatment chemical but they include efforts to limit overtreatment, posttreatment processes to clean the wood surface of chemical deposits, and, in the case of water-based systems,
efforts to immobilize the chemical near the wood surface to reduce the initial flush of chemical when
the wood is immersed in water.

23.6 Summary
Wood is one of the worlds most important renewable natural fibers, and it is used for fuel, housing, railroads, bridges, utility poles, and a host of other applications. Like all materials, wood is also susceptible
to degradation by a variety of living and nonliving agents. Living agents require adequate moisture,
oxygen, and temperature in order to utilize wood, and an important component of wood protection
involves limiting access to one or more of these requirements. The most important requirement is to
limit moisture and most buildings are designed to limit wetting. Where moisture control is not possible,
the wood can be selected from among those species that contain chemicals that are toxic or inhibitory
to degrading organisms or the wood can be supplementally protected. Wood protection can involve
impregnation with pesticides, but it can also utilize compounds that alter wood structure to reduce
the ability to absorb water or make the wood unrecognizable to fungal enzymes. These materials are
typically delivered using combinations of vacuum and pressure to force chemicals into the wood. In
most cases, the treatment is limited to the sapwood, but this limited penetration can be very effective as
long as the protective barrier remains intact. Successful application of protective chemicals can increase
wood service life 8- to 10-fold over untreated wood.

644

Other Natural Materials

Downloaded by [San Jose State University] at 11:34 15 April 2016

References
American Wood Protection Association (AWPA). 2009. Annual Book of Standards, AWPA, Birmingham, AL.
Chou, C.K., J.A. Chandler, and R.D. Preston. 1971. Uptake of metal toxicants by fungal hyphae colonising
CCA-impregnated wood. Wood Science and Technology, 7, 206211.
Evans, P.D., P.D. Thay, and K.J. Schmalzl. 1996. Degradation of wood surfaces during natural weathering.
Effects on lignin and cellulose and on the adhesion of latex primers. Wood Science and Technology,
30, 411422.
Feist, W.C. 1990. Outdoor wood weathering and protection, in R.M. Rowell and R.J. Barbour, Eds.,
Archaeological Wood: Properties, Chemistry, and Preservation, Advances in Chemistry Series 225,
American Chemical Society, Washington, DC, pp. 263298.
Freeman, M.H. and C.R. McIntyre. 2008. A comprehensive review of copper based wood preservatives.
Forest Products Journal, 58(11), 627.
Freeman, M.H., T.F. Shupe, R.P. Vlosky, and H.M. Barnes. 2003. The future of the wood preservation
industry. Forest Products Journal, 53(10), 815.
Dahlgren, S.E. 1974. Kinetics and mechanisms of fixation of Cu-Cr-As wood preservatives. IV. Conversion
reactions during storage. Holzforschung, 28(2), 5861.
Dahlgren, S.E. and W.H. Hartford. 1972a. Kinetics and mechanisms of fixation of Cu-Cr-As wood preservatives. III. Fixation of Tanalith C and comparison of different preservatives. Holzforschung,
26(4), 142149.
Dahlgren, S.E. and W.H. Hartford. 1972b. Kinetics and mechanisms of fixation of Cu-Cr-As wood preservatives. II. Fixation of Boliden K33. Holzforschung, 26(3), 105113.
Dahlgren, S.E. and W.H. Hartford. 1972c. Kinetics and mechanisms of fixation of Cu-Cr-As wood preservatives. I. pH behaviour and general aspects of fixation. Holzforschung, 26(2), 6269.
Duncan, C.G. and F.F. Lombard. 1965. Fungi associated with principal decays in wood products in the
United States. U.S. Forest Service Research Paper WO-4, USDA, Washington, DC.
Graham, R.D. 1973. History of wood preservation, in D.D. Nicholas, Ed., Wood Deterioration and Its
Prevention by Preservative Treatments, Syracuse University Press, Syracuse, NY, pp. 125.
Hale, M.D.C. and R.A. Eaton. 1992. Wood: Decay, Pests and Protection, Chapman & Hall, London, U.K.
Hansen, L.D. and J.H. Klotz. 2005. Carpenter Ants of the United States and Canada, Comstock Publishing
Associates, Cornell University Press, Ithaca, NY.
Harada, H. and W.A. Cote, Jr. 1985. Structure of wood, in T. Higuchi, Ed., Biosynthesis and Biodegradation
of Wood Components, Academic Press, San Diego, CA, pp. 142.
Hillis, W.E. 1962. Wood Extractives, Academic Press, New York.
Hunt, G.M. and G.A. Garratt. 1967. Wood Preservation, McGraw-Hill, New York, 457p.
Lebow, S.T., P.K. Lebow, and D.O. Foster. 2000. Environmental impact of preservative-treated wood
in a wetland boardwalk. Part I. Leaching and environmental accumulation. U.S. Department of
Agriculture Forest Products Laboratory Research Paper FPL-RP-582. Madison, WI, 72p.
Mac Lean, J.D. 1952. Preservative treatment of wood by pressure methods, in USDA Agricultural Handbook
40, Washington, DC, 160p.
Mankowski, M. and J.J. Morrell. 2000. Patterns of fungal attack in wood-plastic composites following
exposure in a soil block test. Wood and Fiber Science, 32, 340345.
Murphy, R.J. and P. Turner. 1989. A vapor phase preservative treatment of manufactured wood-based
board materials. Wood Science and Technology, 23(3), 273279.
Sahle-Demessie, E., K.L. Levien, and J.J. Morrell. 1995. Impregnation of wood with biocides using supercritical fluid carriers, in K.W. Hutchenson and N.R. Foster, Eds., Innovations in Supercritical Fluids:
Science and Technology, American Chemical Society, Washington, DC, pp. 415428.
Scheffer, T.C. 1971. A climate index for estimating potential for decay in wood structures above ground.
Forest Products Journal, 21(10), 2531.

Downloaded by [San Jose State University] at 11:34 15 April 2016

Wood

645

Scheffer, T.C. and E.B. Cowling. 1966. Natural resistance of wood to microbial deterioration. Annual
Review of Phytopathology, 4, 147170.
Scheffer, T.C. and R.M. Lindgren. 1940. Stains of sapwood and sapwood products and their control. U.S.
Department of Agriculture, Technical Bulletin 714, Washington, DC.
Schultz, T.P., D.D. Nicholas, and A.F. Preston. 2007. A brief review of the past, present and future of wood
preservation. Pest Management Science, 63(8), 784788.
Taylor, A.M., B.L. Gartner, and J.J. Morrell. 2002. Heartwood formation and natural durabilityA review.
Wood and Fiber Science, 34(4), 587611.
U.S. Department of Agriculture (USDA). 1980. The biologic and economic assessment of pentachlorophenol, inorganic arsenicals, and creosote, Volume I. Wood preservatives. USDA Technical Bulletin
1658-1. USDA, Washington, DC, 435p.
U.S. Department of Agriculture (USDA). 1996. Wood handbook: Wood as an engineering material. USDA
Forest Service Forest Products Laboratory General Technical Report FPL-GTR-113. Madison,
WI,463p.
Western Wood Preservers Institute. 2006. Best Management Practices: For the Use of Treated Wood in
Aquatic and Other Sensitive Environments, WWPI, Vancouver, WA.
Zabel, R.A. and J.J. Morrell. 1992. Wood Microbiology: Decay and Its Prevention, Academic Press,
SanDiego, CA.

Downloaded by [San Jose State University] at 11:34 15 April 2016

You might also like