ArticlePDF Available

On the changes of pinewood (Pinus sylvestris L.) Chemical composition and ultrastructure during the attack by brown-rot fungi Postia placenta and Coniophora puteana

Authors:

Abstract and Figures

Pine sapwood blocks were exposed to brown-rot fungi Postia placenta and Coniophora puteana for 1, 2, 3, and 4 months. Ion exchange chromatography method was used to determine the changes in sugar content during brown-rot decay. The most remarkable feature was the preferred degradation of mannose both by P. placenta (80.9% weight loss) and C. puteana (77.5% weight loss) in comparison to the control. It can be interpreted as a result of the preferred degradation of the backbone chains of O-acetyl-galacto-glucomannans.For scanning electron microscopy, the wood degradation sequences obtained by the fracture method were used. The fractured surface of cell wall reflected the pattern of degradation of cellulose microfibrils. The surface of secondary wall decayed by P. placenta revealed both smooth and uneven regions throughout the test. C. puteana retained uneven cell wall surface only after the first month, while later the surface became smooth. This suggests that P. placenta tended to degrade cellulose amorphous regions more readily, while C. puteana, possessing full enzyme complement, was able to degrade both amorphous and crystalline regions more readily.The water vapor sorption method was applied for microstructural characteristics of wood. The formation of new pores of size 2.1–9.9 nm occurred during brown-rot decay, whose sizes and volume depend on the fungus culture and exposure time. It is assumed that the appearance of these pores is caused by the destruction of carbohydrates.
Content may be subject to copyright.
International Biodeterioration & Biodegradation 57 (2006) 99– 106
On the changes of pinewood (Pinus sylvestris L.) Chemical
composition and ultrastructure during the attack by brown-rot
fungi Postia placenta and Coniophora puteana
I. Irbe
a,
, B. Andersons
a
, J. Chirkova
a
, U. Kallavus
b
, I. Andersone
a
, O. Faix
c
a
Laboratory of Wood Protection and Emission from Wood Based Products, Latvian State Institute of Wood Chemistry,
27 Dzerbenes Str., LV-1006 Riga, Latvia
b
Centre for Materials Research, Tallinn University of Technology, Estonia
c
Institute of Wood Chemistry and Chemical Technology of Wood, Federal Research Centre for Forestry and Forest Products (BFH), Hamburg, Germany
Received 14 April 2005; accepted 14 December 2005
Available online 15 February 2006
Abstract
Pine sapwood blocks were exposed to brown-rot fungi Postia placenta and Coniophora puteana for 1, 2, 3, and 4 months. Ion exchange
chromatography method was used to determine the changes in sugar content during brown-rot decay. The most remarkable feature was
the preferred degradation of mannose both by P. placenta (80.9% weight loss) and C. puteana (77.5% weight loss) in comparison to the
control. It can be interpreted as a result of the preferred degradation of the backbone chains of O-acetyl-galacto-glucomannans.
For scanning electron microscopy, the wood degradation sequences obtained by the fracture method were used. The fractured surface
of cell wall reflected the pattern of degradation of cellulose microfibrils. The surface of secondary wall decayed by P. placenta revealed
both smooth and uneven regions throughout the test. C. puteana retained uneven cell wall surface only after the first month, while later
the surface became smooth. This suggests that P. placenta tended to degrade cellulose amorphous regions more readily, while C. puteana,
possessing full enzyme complement, was able to degrade both amorphous and crystalline regions more readily.
The water vapor sorption method was applied for microstructural characteristics of wood. The formation of new pores of size
2.1–9.9 nm occurred during brown-rot decay, whose sizes and volume depend on the fungus culture and exposure time. It is assumed that
the appearance of these pores is caused by the destruction of carbohydrates.
r2006 Elsevier Ltd. All rights reserved.
1. Introduction
Brown-rot fungi cause the most considerable decay of
wood in service. These fungi depolymerize and metabolize
wood polysaccharides, leaving behind a brown lignin
residue. Wood polysaccharides are extensively degraded
(Blanchette and Abad, 1988), while lignin is modified or
slightly depolymerized (Jin et al., 1990). Incipient decay of
wood by brown-rot fungi caused measurable strength losses
in wood before measurable weight loss occurs (Wilcox, 1978;
Curling et al., 2002).
The role of small diffusible agent in brown-rot decay is
widely discussed. Several authors (Koenigs, 1974;Illman
et al., 1989;Backa et al., 1992;Goodell et al., 1997) support
the hypothesis that the free radicals produced in the Fenton
reaction (Fe
2+
–H
2
O
2
) are involved in wood degradation
by brown-rot fungi. Extracellular H
2
O
2
is detected in
hyphae of brown-rot fungi, and in the wood cell walls in
the early stages of decay (Kim et al., 2002). This suggests
that H
2
O
2
plays an important role in the early degradation
of cellulose by brown-rot fungi.
This study characterized brown-rotted wood at the
increasing stages of biodeterioration. A better under-
standing of lignocellulose decomposition by brown-rot
fungi could contribute to the application of biochemical
processes in biotechnology and wood decay control.
ARTICLE IN PRESS
www.elsevier.com/locate/ibiod
0964-8305/$ - see front matter r2006 Elsevier Ltd. All rights reserved.
doi:10.1016/j.ibiod.2005.12.002
Corresponding author. Tel.: +371 7545137; fax: 371 7550635.
E-mail address: ilzeirbe@edi.lv (I. Irbe).
2. Materials and methods
2.1. Fungal strains
Two fungal strains were used in this study: Postia placenta (Fr.) Lars
et Lomb. (FPRL 280) and Coniophora puteana (Schum.:Fr.) Karst. (BAM
Ebw.15). Isolates were maintained on malt agar slants at 6 1C. Mycelium was
transferred aseptically to Petri dishes containing 5% malt extract and 3%
Bacto Agar (Ferak, Berlin), pH 6.4, and incubated at 22 1C and 70% RH.
Mycelial plugs were transferred to Kolle flasks for the wood decay test.
2.2. Decay test
Scots pine (Pinus sylvestris L.) sapwood blocks (50 25 15 mm)
were exposed to the fungi for 1, 2, 3 and 4 months at 22 1C and 70% RH.
Test procedures were done according to the European standard EN 113
(1996). Three replicates for the fungus and each test period were taken.
Subsequent to cultivation, blocks were removed from the culture vessels,
brushed free of mycelium and oven dried at 103 721C. The percentage of
weight loss was calculated from the dry weight before and after the test.
2.3. Ion exchange chromatography (IEC)
The carbohydrate composition in brown-rotted wood was determined
by total acidic hydrolysis, followed by the analysis of monomeric sugars
(Uremovic et al., 1994). Samples for quantitative borate IEC were taken
from wood blocks decayed by the fungi for 2 and 4 months. The samples
were taken from the middle part of the wood block after its cutting in a
cross-direction.
Wood material was ground and then subjected to acid hydrolysis. In
the primary hydrolysis, 200 mg of dried sawdust was hydrolyzed with 72%
sulfuric acid at 30 1C for 1 h. The secondary hydrolysis was performed
after dilution with water to a 4% sulfuric acid concentration by
autoclaving at 120 1C for 40 min. Then the hydrolysate was filtered on a
glass filter to determine its lignin part. The acidic filtrate containing sugars
was directly applied to a column, filled with anion-exchanger resin. The
mobile phase was 0.475 M potassium-borate buffer, pH 9.2. 2,2-
bicinchoninate reagent for reducing sugars was used for detection of
sugars in the column eluate.
2.4. Scanning electron microscopy (SEM)
Two sequences of wood degraded by P. placenta and C. puteana for 4
months were obtained from the wood block lower surface close to the
substrate.
Wood blocks were sawn into two parts (1/3 and 2/3), then the larger
piece was split with a knife and broken to get the fractured surface. After
getting a proper sample, it was attached to the specimen stub with
conductive glue, coated with gold and examined with a JEOL scanning
electron microscope JSM 840A at 15 kV. All micrographs taken at the
same magnification. In this study, only latewood cells were used, as thicker
cell walls showed more distinct results.
2.5. Water vapor sorption (WVS)
Wood samples after fungal degradation were dried at 103 1C, milled,
and a fraction of size 0.5–1.0 mm was taken for measurement of the
isotherms. Water vapor sorption–desorption isotherms were measured on
a vacuum balance with quartz spirals as the sensible element (the
sensitivity was 1.5–2.0 mm/mg) at the temperature 22 70.1 1C.
The elongation of the spirals was measured by a horizontal microscope
with the accuracy 0.005 mm. The time of reaching the equilibrium in each
point of the isotherm was 20–24 h.
The isotherms were analyzed by the comparative method in combina-
tion with the BET method (Chirkova et al., 2004). The following
characteristics of the samples’ microstructure were determined: accessible
specific surface (A,m
2
/g), mass and surface concentrations of hydrophilic
centers (a
m
, mmol/g and a, groups/nm
2
, respectively), volume of the
sorption space (W,cm
3
/g) and average pore size ðDaver ¼ð4WÞ=AÞ.
The distributions of the pores’ volume in terms of their sizes
½dW=dD¼fðDÞwere calculated from the desorption branch of the
isotherms. The D-values were calculated from the Kelvin equation:
lnðP=P0Þ¼gv=DRT,
where gis the surface tension of the sorbate; vthe molar volume of the
liquid sorbate; Rthe gas constant; and Tis the temperature.
3. Results and discussion
3.1. Ion exchange chromatography
The composition of control specimens truly reflected the
composition of softwood polysaccarides. They consist of
approx. 43% cellulose, 10–15% arabino-4-O-methyl-
glucurono-xylan (Xyl/GluA/L-Ara ¼10/2/1.3), 5–10% O-
acetyl-galacto-glucomannan (Man/Glu/Gal ¼3/1/1), and
27% lignin (Timell, 1967).
Glucose content in the control sample after hydrolysis
was the highest (49.1 abs.%) followed by mannose, xylose,
galactose, and arabinose (12.9, 5.7, 1.7, and 1.2 abs.%,
respectively) (Table 1). The number calculated by differ-
ence to 100 (29.4% for the control) was mostly due to the
lignin and resin content. Experimental hydrolysis losses
could also contribute to this figure.
The weight loss of the samples was an important
measure of the fungi’s bioactivity being after 4 month of
exposure 47% (P. placenta) and 40% (C. puteana).
Conclusive is the change of the sugar composition,
expressed both in abs.% and rel.%, in the hydrolysates
of degraded samples. Table 1 reveals that the sum of the
detectable sugars (Sabs.%) is lower than the residual
amount of solid wood left after degradation (100—weight
loss). After 4 month of exposure to P. placenta, for
example, 29.9 g sugar was detectable in 53 g of residual
wood. In the case of C. puteana 40.6 g sugar was found in
60 g of degraded wood (Fig. 1).
In this context, the difference between the corresponding
values can be explained by the presence of non-hydro-
lysable wood moieties. These consist mostly of lignin. In
the case of pine the presence of resins cannot be excluded.
Carbohydrate conversion products, as a result of metabo-
lism, may also be difficult to determine by routine IEC
sugar analysis. Moreover, hydrolysis losses may occur,
though to a small extent.
Table 1 also give more details of sugar composition; the
most remarkable feature is the preferred degradation of
mannose both by P. placenta (80.9% weight loss) and C.
puteana (77.5% weight loss) in comparison to the control.
By contrast, in the case of the glucose, enrichment can be
observed. For example, after 4 month 78.5 rel.% (P.
placenta) and 81.8% (C. puteana) glucose was found in the
hydrolysate, while in the control 69.5 rel.% glucose was
present. In absolute values, the amount of glucose changed
ARTICLE IN PRESS
I. Irbe et al. / International Biodeterioration & Biodegradation 57 (2006) 99–106100
dramatically compared to other sugars, i.e. totally glucose
was consumed in the highest amount. In 4 months, P.
placenta degraded glucose from 49.1% (initial) to 23.5%
and C. puteana to 33.2%.
These observations can safely be interpreted as a result
of the preferred degradation of the backbone chains of O-
acetyl-galacto-glucomannans. This polymer is probably
associated with less crystalline parts of the microfibril.
Highley (1987) had demonstrated that glucomannans of
Pinus monticola are more easily metabolized by brown-rot
than xylans. In a study of southern pine degradation by
brown-rot was shown that initial strength loss (up to 40%)
without measurable weight loss corresponded to loss in the
galactan and arabinan components (Curling et al., 2002).
Subsequent strength loss (greater than 40%) and significant
weight loss (5–20%) were associated with the loss of the
mannan and xylan components. Major loss of glucan
component (representing cellulose) was achieved at 475%
strength loss and 420% weight loss. Our results also
showed (Table 1) a slight loss of glucose at the initial stage
of wood decay. 12.6% (P. placenta) and 8.9% (C. puteana)
of glucose were lost in 2 months at 20% weight loss, while
hemicellulose sugars in this period were lost at significantly
higher amounts. This can be attributed to the low
accessibility of (crystalline) cellulose to the fungal enzyme
system.
3.2. Scanning electron microscopy (SEM)
In this study the wood degradation sequences obtained
by the fracture method were included as they provided
ARTICLE IN PRESS
70.6
29.4
55.2
44.8
20
29.9
70.1
47
56.5
43.5
21
40.6
59.4
40
0%
20%
40%
60%
80%
100%
control P. placenta
2 months
P. placenta
4 months
C. puteana
2 months
C. puteana
4 months
carbohydrates lignin weight losses
Fig. 1. Relation among weight losses, non-hydrolysable (ligin) and hydrolysable (carbohydrates) wood components after decay by P. placenta and
C. puteana.
Table 1
Weight losses and changes in wood components after pinewood degradation by brown-rot fungi P. placenta and C. puteana
Fungus Exposure
(month)
Weight
loss (%)
a
Glucose (%) Mannose (%) Xylose (%) Galactose (%) Arabinose (%) SS100 minus
Abs. Rel. Loss Abs. Rel. Loss Abs. Rel. Loss Abs. Rel. Loss Abs. Rel. Loss Abs.
(%)
Rel.
(%)
S
Abs. (%)
Control 49.1 69.5 12.9 18.3 5.7 8.0 1.7 2.4 1.2 1.8 70.6 100 29.4
P. placenta 1672
22071 42.9 77.7 12.6 6.6 12.0 48.7 3.9 7.1 31.2 0.9 1.6 48.5 0.9 1.7 25.0 55.2 100 44.8
33671
44774 23.5 78.5 52.2 2.5 8.3 80.9 2.7 8.9 53.1 0.7 2.3 59.2 0.6 2.0 51.6 29.9 100 70.1
C. puteana 11272
22171 44.7 78.9 8.9 6.1 10.8 53.1 4.0 7.1 23.0 0.9 1.5 49.7 1.0 1.7 23.4 56.5 100 43.5
32872
44071 33.2 81.8 32.3 2.9 7.2 77.5 3.0 7.3 47.4 0.7 1.8 56.2 0.8 1.9 38.7 40.6 100 59.4
a
Mean (3) 7standard deviation.
I. Irbe et al. / International Biodeterioration & Biodegradation 57 (2006) 99–106 101
more detailed view on wood cell structure in comparison
with the cutting method.
Degradation sequence of P. placenta attacked wood is
shown in Fig. 2. The fractured surface of cell wall gradually
turned smoother but the differences between fractures after
1 and 3 months were almost unnoticed. As the degradation
of cellulose turned deeper (the mass loss increased) the
surface became smoother, with glass-like appearance. After
4 months, the surface was less uneven than after 1–3
months. By the end of test, a visible boundary between the
cell wall and the compound middle lamellae (CML) was
still observed.
Fig. 3 shows the sequence of wood attacked by C.
puteana. After 1 month the fracture surface was rough,
with deep helical cracks and robust sliced surface. Starting
from the 2nd to the 4th month the cell wall broke evenly
showing smooth structure. The S
3
layer of the secondary
wall had already cracked after 2 months of attack leaving
deep canyons along the lumen surface. Morphological
differences between the secondary wall and the CML were
well distinguished after 1 month of degradation. The cell
wall surface revealed a sliced structure across to the fiber
axis, while the CML showed smooth structure. The
difference between both layers disappeared in the later
degradation periods, when the secondary wall looked like
the CML. The cell wall thinned in a similar way when
degraded by white-rot fungi, which remove all cell wall
components causing a widespread thinning of the wall
(Eriksson et al., 1990).
The micrographs showed difference in wood degradation
patterns between brown-rot species under this study.
According to the IEC (Table 1), C. puteana could be
regarded as less active in consuming cellulose, while SEM
showed opposite results i.e., smooth breaking surface
already in early decay stages. The smooth feature observed
in the fractured S
2
layer can be explained (Encinas et al.,
1998) as a consequence of the effects on the degree of
depolymerization of cellulose chains by enzymes. It is
supposed that C. puteana caused rapid depolymerization of
cellulose chains throughout the cell wall without a
considerable loss of cellulose.
Weight loss in the case of both fungi after 4 months
fluctuated in the range of 40–47% with the higher value of
P. placenta, while SEM micrographs illustrated uneven
wood breakdown by the fungus throughout the test.
According to Highley (1977),P. placenta contains en-
doglucanases and glycosidases, but not exoglucanases,
while Coniophoroid brown-rots produce the full enzyme
complement (Highley, 1988). Endoglucanases are supposed
to attack in the middle of the more disordered regions of
cellulose, and exoglucanases the crystalline areas at the
opposite chain ends. A typical endoglucanase cleaves the
cellulose chains, resulting in a rapid decrease in the degree
of polymerization (DP), while exoglucanases decrease the
substrate DP very slowly. The endo–exo synergy enhances
the activity of the individual enzymes (Teeri, 1997). It is
supposed that P. placenta tended to degrade cellulose
amorphous regions more readily, while the crystalline ones
remained less damaged. This could affect the wood fracture
and reveal the cell wall surface as uneven. C. puteana with
the endo–exo enzymes was able to degrade both amor-
phous and crystalline regions more readily, that appeared
ARTICLE IN PRESS
Fig. 2. Degradation sequence of wood blocks decayed by P. placenta in 4 months.
Fig. 3. Degradation sequence of wood blocks decayed by C. puteana in 4 months.
I. Irbe et al. / International Biodeterioration & Biodegradation 57 (2006) 99–106102
in micrographs as a smooth fractured surface. The
fractured surface of wood cell reflected the pattern of
degradation of cellulose microfibrils.
SEM micrographs showed that the S
3
layer remained
sound after P. placenta attack, but was altered in some
locations after C. puteana decay (Fig. 3). Endo- and
exoglucanases of C. puteana probably more intensively
affected the S
3
layer causing erosion and thinning of the
cell wall. Degradation of the S
2
layer by brown-rot fungi
without causing noticeable loss of the S
3
has been
previously reported by Highley and Murmanis (1985).It
is because of a more compact layer of residual lignin in the
S
3
layer. The alterations in the S
3
could be explained by the
differences in enzymatic systems between both brown-rot
species.
CML in brown-rotted wood revealed smooth structure
from the beginning to the end of test. In P. placenta-
decayed wood, a visible boundary between the secondary
wall and CML was observed by the end of test. In C.
puteana-attacked wood, the difference between both layers
had disappeared by the second month of degradation,
when the secondary wall looked like the CML. We assume
that the smooth fractured surface of the CML was because
of a high amount of lignin. The cell wall ultrastructure of
brown-rotted wood studied by TEM (Eriksson et al., 1990)
provided evidence of deposits of unhydrolysed lignin
throughout the cell wall, with the greatest amount in the
CML. It is known (Fengel and Wegener, 1989) that the
CML of softwoods contains high amount of lignin, around
60%, in comparison with the secondary wall where it is
around 27%. Cellulose comprises around 14% in CML,
while in secondary wall it is around 60%.
3.3. Water vapor sorption
Fig. 4 shows the isotherms of sorption–desorption of
water vapors of wood samples after a 2-month exposure
with the fungi C. puteana and P. placenta in comparison
with the corresponding isotherms of sound wood control.
For all isotherms, the presence of hysteresis in the whole
range of P/P
0
is typical, which is connected with a partial
inclusion of the sorbate in the structure of non-rigid
sorbent upon desorption.
The hysteresis loop width depends on the strength of
interstructural bonds of the sorbent and grows with their
increase. It narrows on all degraded samples after 1 month
of exposure (weakening of bonds after the removal of
hemicelluloses). With increasing exposure time, the P.
placenta hysteresis loop grows until 4 months of exposure
owing to the formation of new interstructural bonds after
the sample’s biodegradation. In the case of C. puteana, the
widest hysteresis loop occurs after 3 months of exposure.
Fig. 5 shows the change in the structural characteristics
of wood depending on the exposure time. The initial (after
1 month) decrease of specific surface (A) is connected with
the removal of hemicelluloses, which, owing to a low MW,
ensures a higher swelling in the initial wood. In this case,
the concentration of hydrophilic centers (a) grows. These
centers are released upon the removal of low-molecular
carbohydrates and do not manage to compensate at the
expense of re-structuring. The further decrease in the A
value occurs owing to the decrease of MW of carbohy-
drates due to degradation, as well as the enrichment of
samples in lignin, whose swelling in water vapors is lower
in comparison with that of carbohydrates.
ARTICLE IN PRESS
Control
0
50
100
150
200
250
300
0 0.2 0.4 0.6 0.8 1
0 0.2 0.4 0.6 0.8 1
0 0.2 0.4 0.6 0.8 1
a (mg/g)
P. placenta
0
50
100
150
200
250
300
a (mg/g)
C. puteana
-50
0
50
100
150
200
250
300
P/Po
a (mg/g)
Fig. 4. Water vapor sorption–desorption isotherms of sound wood and
wood decayed by P. placenta and C. puteana for 2 months (sorption—
circles; desorption—cubes).
I. Irbe et al. / International Biodeterioration & Biodegradation 57 (2006) 99–106 103
The mass concentration of hydrophilic centers (a
m
) after
1 month of exposure on both cultures grew, since the
increase in surface concentration (a) prevailed over the
decrease in the specific surface. Hereafter (up to 3 months
of exposure), the a
m
value decreased owing to the decrease
of both Aand a. After a 4-month exposure on fungi, the
decrease of the Avalue on P. placenta slowed down, and
even grew somewhat on C. puteana in the case of a
simultaneous increase of the surface concentration of the
hydrophilic centers, which resulted in an increase in a
m
.
The average pore size for the swollen wood grew with
increasing biodeterioration time (Fig. 5), reaching the
highest values (4.5 nm) after 3 months of exposure. It
should be mentioned that the samples investigated were
subjected to drying at 103 1C after the exposure. Such a
treatment caused a partial collapse of the structure, which
resulted in a partial loss of information during the analysis
of the sorption branch of the isotherm. However, in the
process of saturating with water vapors, the wood structure
relaxed, and the desorption branch of the isotherm could
be regarded as more equilibrium.
From the desorption branches of the isotherms,the curves
of distribution in volume sizes were plotted for the pores in
size range 3–200 nm (Fig. 6). Table 2 lists the sizes and
volumes of such pores in biodeteriorated wood. These data
show that new pores (D¼2:1–2.5 nm) appeared in the
samples exposed to P. placenta after the first month of the
experiment. With time, the size and volume of these pores
grew. These values were close to those found by Flournoy
et al. (1991), who investigated the porous structure of moist
wood after the decay by P. placenta for 1 month by the
method of sorption from aqueous solutions of substances
with different molecular sizes.
In contrast to P. placenta, the degrading action of C.
puteana manifested itself in a different way. new pores in
the specified size range appeared only after a 3-month
exposure, and their size and volume were less than those of
P. placenta.
It is of interest to elucidate the reasons for the
appearance of these pores and the distinctions in the
behavior of the two fungi. Flournoy et al. (1991) assigned
the newly developed pores to the sizes of the active agent of
the fungus. However, even the cell wall of sound wood in
the swollen state contains a great volume of rather wide
pores. The average pore size for intact pinewood was
3.8 nm. Obviously, such pores, being absent in dry wood,
are accessible not only to low-molecular substances, but
also to rather large molecules. Besides, the actual distribu-
tion of pore volumes in their sizes should be taken into
account, namely, the wood samples had pores, whose size
exceeds the average one.
At the same time, all known enzymes exceed the
micropore size of the intact structure of the wood matrix,
while brown-rot fungi employ a low molecular weight
decay system and rapidly depolymerize cellulose (Goodell
et al., 2003).
The formation of new pores ranged from 2.1 to 9.9 nm
(Table 2) depends on the fungus and exposure time.
It is assumed that the appearance of small pores is
connected with the destruction of cellulose and the
‘‘imperfection’’ of the wall structure. The essential distinc-
tion between the two fungi under study is, as has been
mentioned above, in the different nature of enzymes,
namely, endo- and exoglucanases. Probably, the degrada-
tion of the cell wood attacted by C. puteana occurs firstly
mainly on the wall surface, but in the case of P. placenta
the destruction of cellulose even after 1 month of exposure
ARTICLE IN PRESS
100
150
200
250
300
350
01234
01234
01234
01234
A, m2/ g
1
1.5
2
2.5
3
3.5
am, mM/g
3
4
5
6
7
8
α, groups/nm2
2
2.5
3
3.5
4
4.5
5
Time (months)
Daver, nm
P. placenta C. puteana
Fig. 5. Microstructural characteristics (specific surface A, mass and
surface concentrations of hydrophilic centers a
m
and a, average pore size
D
aver
) of pinewood depending on the exposure time on P. placenta and C.
puteana.
I. Irbe et al. / International Biodeterioration & Biodegradation 57 (2006) 99–106104
proceeds deep inside the wall. This assumption is indirectly
confirmed by microphotographs (Figs. 2 and 3), as well as
the higher activity of P. placenta in comparison with C.
puteana.
Acknowledgement
The authors gratefully acknowledge Dr. J. Puls (Institute
of Wood Chemistry and Chemical Wood Technology,
Federal Research Centre for Forestry and Forest Products,
Hamburg) for the fruitful discussion of the polysaccharides
part of biodegraded wood.
References
Backa, S., Grirer, J., Reitberger, T., Nilsson, T., 1992. Hydroxyl radical
activity in brown-rot fungi by a new chemiluminiscence method.
Holzforschung 46, 61–62.
Blanchette, R.A., Abad, A.R., 1988. Ultrastructural localization of
hemicellulose in birch wood (Betula papyrifera) decayed by brown
and white rot fungi. Holzforschung 42, 393–398.
Chirkova, J., Andersons, B., Andersone, I., 2004. Determination of
standard isotherms of the sorption of some vapors with cellulose.
Journal of Colloid and Interface Science 276, 284–289.
Curling, S.F., Clausen, C.A., Winandy, J.E., 2002. Relationships between
mechanical properties, weight loss, and chemical composition of wood
during incipient brown-rot decay. Forest Products Journal 52 (7/8),
34–39.
Encinas, O., Henningsson, B., Daniel, G., 1998. Changes in toughness and
fracture characteristics of wood attacked by the blue stain fungus
Lasiodiplodia theobromae. Holzforschung 52, 82–88.
Eriksson, K.E., Blanchette, R.A., Ander, P., 1990. Microbial and
Enzymatic Degradation of Wood and Wood Components. Springer,
Berlin 407pp.
European Standard EN 113, 1996. Wood preservatives. Method of Test
for Determining the Protective Effectiveness Against Wood Destroy-
ing Basidiomycetes. CEN, Bruessel.
Fengel, D., Wegener, G., 1989. Wood (Chemistry, Ultrastructure,
Reactions). Walter de Gruyter, Berlin, New York 613pp.
Flournoy, D.S., Kirk, T.K., Highley, T.L., 1991. Wood decay by brown-
rot fungi: changes in pore structure and cell wall volume. Holz-
forschung 45, 383–388.
Goodell, B., Jellison, J., Liu, J., Daniel, G., Paszczynski, A., Fekete, F.,
Krishnamurthy, S., Jun, L., Xu, G., 1997. Low molecular weight
chelators and phenolic compounds isolated from wood decay fungi
and their role in the fungal biodegradation of wood. Journal of
Biotechnology 53, 133–162.
Goodell, B., Nicholas, D.D., Schultz, T.P., 2003. Wood Deterioration and
Preservation. American Chemical Society (ACS), Washington, DC
465pp.
Highley, T.L., 1977. Degradation of cellulose by culture filtrates of Poria
placenta. Material und Organismen 12, 161–174.
Highley, T.L., 1987. Changes in chemical components of hardwood and
softwood by brown-rot fungi. Material und Organismen 21, 39–45.
Highley, T.L., 1988. Cellulolytic activity of brown-rot and white-rot fungi
on solid media. Holzforschung 42, 211–216.
Highley, T.L., Murmanis, L., 1985. Micromorphology of degradation in
western hemlock and sweetgum by the brown-rot fungus P. placenta.
Holzforschung 39, 73–78.
Illman, B.L., Meinholtz, D.C., Highley, T.L., 1989. Oxygen free radical
detection in wood colonized by the brown-rot fungus Postia placenta.
Biodeterioration Research 2, 497–509.
Jin, L., Schultz, T.P., Nicholas, D.D., 1990. Structural characterisation of
brown-rotted lignin. Holzforschung 44, 133–138.
Kim, Y.S., Wi, S.G., Lee, K.H., Singh, A.P., 2002. Cytochemical
localization of hydrogen peroxide production during wood decay by
brown-rot fungi Tyromyces palustris and Coniophora puteana. Holz-
forschung 56, 7–12.
Koenigs, J.W., 1974. Production of hydrogen peroxide by wood
rotting fungi in wood and its correlation with weight loss,
ARTICLE IN PRESS
0
0.01
0.02
0.03
0.04
0.05
0.06
0.07
0.08
0.09
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5
r, nm
dw/dr
P. placenta, 2 months P. placenta, 4 months
C. puteana, 2 months C. puteana, 4 months
Fig. 6. The distribution of pores by specific size in brown-rotted wood.
Table 2
Characteristics of the porous structure of destructed wood samples
Fungus Time Drange (nm) Volume of pores
(months) in the range (cm
3
/g)
Control —
P. placenta 1 2.1–2.5 0.015
2 3.0–83 0.065
3 3.1–9.9 0.055
4 3.1–9.9 0.070
C. puteana 1— —
2— —
3 2.3–3.6 0.020
4 2.7–8.3 0.035
I. Irbe et al. / International Biodeterioration & Biodegradation 57 (2006) 99–106 105
depolimerization and pH changes. Archives of Microbiology 99,
129–145.
Teeri, T.T., 1997. Crystalline cellulose degradation: new insight into the
function of cellobiohydrolases. Trends in Biotechnology 15, 160–167.
Timell, T.E., 1967. Recent progress in the chemistry of wood hemi-
celluloses. Wood Science and Technology 1, 45–70.
Uremovic, A., Glawischnig, T.D., Schuseil, J., Saake, B., Borchmann, A.,
Herrmann, A., Puls, J., 1994. Chromatographische Untersuchungen
zur quantitativen Bestimmung der Holzzucher. Holz Roh-Werkstoff
52, 347–354.
Wilcox, W.W., 1978. Review of literature on the effect of early stages of
decay on wood strength. Wood Fiber 9 (4), 252–257.
ARTICLE IN PRESS
I. Irbe et al. / International Biodeterioration & Biodegradation 57 (2006) 99–106106
... It was worth mentioning that a similar situation was all the ligninase activities reached their maximum levels after middle pretreatment stage whether in white rot or brown rot groups. This could be due to the depolymerization of carbohydrates and the increase porosity of wood cell walls at the initial fungal pretreating process due to the high hydrolase activities, which was conducive to the disruption of lignin in the advanced period (Aguiar et al., 2013;Arantes and Goodell, 2014;Hunt et al., 2004;Irbe et al., 2006). ...
... X-ray diffraction was particularly used to evaluate the changes of crystallinity in various fungal treatment samples and the results were shown in Table 2 and Fig. 2. In white rot fungus T. versicolor group, the crystallinities decreased through the pretreating stages, particularly at 60 days with the decrease by 5.30%, indicating that the crystal areas of cellulose were mainly attacked at middle process. The main reason was that endoglucanase randomly destroyed the β-1,4 glycosidic bond, and depolymerized crystalline cellulose to form amorphous cellulose, then exdoglucanase could cut two glucose molecules from the end of the sugar chain and destroy the cellulose intensively in the advance stage (Irbe et al., 2006). In G. trabeum and R. placenta groups, the crystallinities of samples continued to decline, with the decreased crystallinities by 4.96% in G. trabeum and 12.76% in R. placenta, indicating that it could cut cellulose macromolecular chain (Encinas et al., 1998) and decrease the crystallinities. ...
... R. placenta and C. puteana additionally utilize a non-enzymatic degradative mechanism based on the production of hydroxyl radicals through the Fenton reaction for the initial depolymerization of cell wall components in order to access carbohydrates prior to the enzymatic wood decay (Arantes and Goodell 2014;Ringman et al. 2014;Zhang et al. 2016). However, differences between these two brown-rot species have been noted in decay tests against wood (Irbe et al. 2006;Metsä-Kortelainen and Viitanen 2009) and studies regarding the gene expression (Sista Kameshwar and Qin 2020). Metsä-Kortelainen and Viitanen (2009) observed that R. placenta caused a relatively high mass loss to thermally treated wood when compared to C. puteana. ...
... Metsä-Kortelainen and Viitanen (2009) observed that R. placenta caused a relatively high mass loss to thermally treated wood when compared to C. puteana. Additionally, Irbe et al. (2006) noted the more rapid formation of new pores by R. placenta. ...
Article
Full-text available
Pressurized hot water extraction (HWE) treatment has the benefit of simultaneous extraction of hemicellulose-based carbohydrates and modification of the solid phase, but it does not drastically improve wood durability. However, removing hemicelluloses from the wood by HWE treatment creates water-filled spaces in the cell walls which could be filled with modification agent in order to improve the properties of the wood. Without drying, modification agent can be added into the saturated wood via diffusion. The esterification of wood with citric acid (CA) improves resistance to biological deterioration but increases brittleness. However, combining CA esterification with additional chemicals that form links with CA can mitigate brittleness. This study investigated esterification as a method for modifying HWE treated wood. HWE treatment with CA solution (4% w/v) was applied at 120 °C for 3 h to Scots pine (Pinus sylvestris L.) sapwood specimens. The specimens were further modified by diffusion with CA and starch derivatives followed by curing. The applied method changed the moisture properties and chemical composition of the wood. The results showed successful wood bulking. The investigated method slightly improved decay resistance to Coniophora puteana and Trametes versicolor but did not change resistance to Rhodonia placenta.
... This phenomenon resulted from enhanced reaction rates due to higher water activity (Panchariya et al. 2001). Moreover, the width of the hysteresis loop depends on internal bonding between and among the cell wall polymers (Irbe et al. 2006). Higher measurement temperatures also reduced the sorption hysteresis of all wood samples, especially in the high humidity range, with a maximum hysteresis seen at 60-70% RH. ...
Article
Full-text available
Social or community-managed forestry has been promoted as an inclusive way of mitigating climate change, under the UNFCCC’s REDD+ programme (Reducing Emissions from Deforestation and forest Degradation). Fast-growing wood species are little used as a community forestry resource in South-East Asia, although their potential could help to meet the surge in demand for sawn timber. In 2016, the Government of Indonesia aimed to hand over concession rights to 12.7 million hectares for community forestry, and support at the national level is currently strong. Although there is high demand for commercial plantation-grown species such as sengon (Paraserianthes falcataria L. Nielsen) and jabon (Anthocephalus cadamba Roxb.), the potential of other species, such as acacia (Acacia mangium Willd.), deserves to be explored. Paulownia (Paulownia tomentosa (Thunb.) Steud.) was also used in this study. Physiologically, fast-growing wood species differ from long-rotation hardwoods in their quality, i.e. their resistance to biodeterioration. In this study, the natural durability of the aforementioned fast-growing species was investigated by laboratory testing, using basidiomycete monocultures. Wood specimens were incubated with brown rot (Coniophora puteana) and white rot (Trametes versicolor) fungi for 16 weeks. Parameters such as mass loss, surface hardness, sorption properties, and anatomical characteristics after exposure to the fungi were determined. Sorption measurement times at high temperature were faster with lower hysteresis. Different relative humidity levels affected the changes in the total vessel area. The fast-growing wood species A. cadamba, P. falcataria, and P. tomentosa were classified as not very durable to non-durable, except for A. mangium, which was classified as durable. The data for specimens exposed to C. puteana were found to be highly variable. The remaining axial hardness of wood specimens incubated with T. versicolor was lower compared to C. Putanea. EDX observation (energy-dispersive X-ray spectroscopy) showed K as the major cation in decayed specimens.
... The width of the hysteresis loop in wood depends on internal bonding between individual cell wall polymers. With the increased number of bonds, the loop increases [32]. In the case of treatments on softwood, pine-AL exhibited a stable performance across the entire moisture range, while pine-LNPs showed stability from 75% relative humidity (RH) onwards. ...
Article
Full-text available
Recent advancements in wood modification aim to enhance the inherent qualities of this versatile biological material, which includes renewability, ease of processing, and thermal insulation. This study focuses on evaluating the effectiveness of lignin as a protective agent for less durable wood species, namely, Pinus nigra and Fagus sylvatica L. The impregnation of wood with three various forms of lignin, such as kraft lignin, acetylated kraft lignin, and lignin nanoparticles, was carried out using the vacuum technique at room conditions. The results showed that the treatments significantly improve the hydrophobicity of wood surfaces, particularly in pine wood, and provide protection against UV ageing. Additionally, the treatments contributed to the stabilisation of moisture content at different humidity levels. Although slight colour variations were observed, their impact on the visual appearance was minimal, and the thermal analysis confirmed enhanced thermal properties. Additionally, plasma treatment further enhanced hydrophobicity after treatments, offering potential benefits in terms of moisture resistance and durability. The findings of this study highlight the promising effects of lignin-based treatments on wood properties, providing sustainable solutions for wood protection in various sectors. However, further optimisation is needed to fully explore the potential of lignin and lignin nanoparticles.
... C. puteana exhibited the lowest weight losses in all cases, and P. placenta was found to be more aggressive than C. puteana attacks. Cell wall degradation by C. puteana occurs firstly mainly on the wall surface, but in the case of P. placenta, cellulose destruction occurs deep inside the wall after the first month of exposure [35]. This assumption may indicate the higher activity of P. placenta in comparison with C. puteana. ...
Article
Full-text available
In this study, spruce wood was modifed by ring-opening polymerization of ε-caprolactone to graft poly (ε-caprolactone) into wood cell wall by impregnation of 30%, 50%, and 70% monomer concentrations and further polymerization in DMF solution. The biodegradability of the modifed wood by the wood-decaying fungi was investigated by means of weight losses, and the chemical and morphological background of the degradation process was analyzed through FTIR and SEM analysis, respectively. For this purpose, modifed samples were exposed to brown rot fungi Coniophora puteana and Postia placenta, and white rot fungi Trametes versicolor and Pleurotus ostreatus attacks to determine the optimum concentration level of ε-caprolactone monomer for sufcient decay resistance on media inoculated with malt-extract agar and soil, according to principles of CEN EN 113 and ASTM D 1413, respectively. A leaching test was conducted in order to evaluate any loss in efectiveness in decay resistance. Results showed that all concentrations of PCL exhibited superior decay resistance in samples after the decay test was conducted on agar media. However, modifed samples gave high weight losses in soil contact decay testing. P. placenta and P. ostreatus attacks were found to be more aggressive in modifed samples than other fungi attacks. 70% ε-caprolactone concentration was found more efcacious in suppressing brown rot fungi attacks than lower concentrations, whilst lower concentration levels were found to be more efcacious in suppressing white rot fungi attacks than 70% concentration level. SEM and FTIR fndings proved that weight losses were due to both cell wall degradations and polymer digestion by fungal enzymes. SEM study revealed that cell wall modifcation inhibits the consumption of cell wall polymers compared to controls.
... The FPDS exhibited a vast increase in the surface area, reaching 903 m 2 g − 1 . A larger surface area indicates that the FPDS has a more porous structure (Irbe et al., 2006). The total pore volume, micropore volume and mesopore volume of FPDS were 0.932 cm 3 g − 1 , 0.386 cm 3 g − 1 , and 0.191 cm 3 g − 1 , respectively. ...
Article
Full-text available
In recent years, enzymatic remediation/biocatalysis has gained prominence for the bioremediation of recalcitrant chemicals. Laccase is one of the commonly investigated enzymes for bioremediation applications. There is a growing interest in immobilizing this enzyme onto adsorbents for achieving high pollutant removal through simultaneous adsorption and biodegradation. Due to the influence of the biomolecule-support interface on laccase activity and stability, it is crucial to functionalize the solid carrier prior to immobilization. Date stones (PDS), as an eco-friendly, low-cost, and effective natural adsorbent, was utilized as a carrier for laccase (fungus Trametes versicolor). After activating PDS through chemical treatments, the surface area increased by thirty-six-fold, and carbonyl groups became more prominent. Batch experiments were carried out for ketoprofen and aspirin biodegradation in aqueous solutions. After six cycles, the laccase maintained 54% of its original activity confirmed by oxidation tests of 2,2′-azino-bis (3-ethylbenzothiazoline-6-sulphonic acid) (ABTS). In addition, the storage, pH, and thermal stability of immobilized laccase on functionalized date stone (LFPDS) were found to be superior to that of free laccase, demonstrating its potential for ongoing applications. In the aqueous batch mode, this immobilized laccase system was used to degrade 25 mg L⁻¹ of ketoprofen and aspirin, resulting in almost complete removal within 4 h of treatment. This study reveals that agricultural wastes such as date stone can successfully be valorized through simple activation techniques, and the final product can be used as an adsorbent and substrate for immobilization enzyme. The high efficiency of the LFPDS in removing ketoprofen and aspirin highlights the potential of this technology for removing pharmaceuticals and merits its continued development.
Article
Full-text available
Hygroscopicity is one of the most important properties of wood and plays a decisive role in its dimensional stability. In this context, conservation plans for waterlogged archaeological wood (WAW) and relevant waterlogged artefacts must be created. The size of the sample required for a moisture sorption assessment may affect the results for (and thus the perception of) the hygroscopicity of a testing artefact. Herein, to investigate the effects of the sample size on the hygroscopicity of WAW as measured via dynamic vapour sorption (DVS), typical WAW and recent (i.e., sound) wood are processed into four differently sized samples, ranging in thickness from 200 mesh to millimetre. The equilibrium moisture contents (EMCs) of the wood samples are simultaneously measured using simultaneous DVS. The sorption isotherms show that the EMC values of the recent wood at each relative humidity increase as the sample size decreases, with the superfine powder sample achieving the highest EMC of all of the recent samples. Although the WAW has a higher EMC than that of recent wood, the effect of the size of the WAW sample on its hygroscopic properties is surprisingly not as pronounced as that for the recent wood. In addition, the hysteresis between the samples of different sizes of the archaeological wood is significantly smaller than that for the reference samples. Furthermore, regarding the standard deviations of the parameters obtained from the Guggenheim Anderson de Boer and Hailwood–Horrobin models, the values for WAW are all much smaller than those for the reference wood. This further verifies the disappearance of the size effect of the hygroscopicity for WAW.
Article
Full-text available
Wood fiber mat-reinforced composite (WFMRC) is a novel type of engineering wood-based polymer composite for decorative and building applications. Fungal decay strongly influences the service life of wood composites in outdoor environments. In this study, fungal durability tests were conducted to investigate how white- and brown-rot fungi (Trametes versicolor and Gloeophyllum trabeum, respectively) affect the chemical composition, crystallinity, and morphology of WFMRCs made from two different wood species (poplar and larch). The poplar WFMRC lost more mass (18.3% and 18.4% by T. versicolor and G. trabeum, respectively) than the decayed larch WFMRC (11.4% and 18.4% by T. versicolor and G. trabeum, respectively) after 12 weeks of fungal exposure. Chemical analysis and Fourier-transform infrared spectroscopy measurement revealed the degradation of holocellulose of the WFMRCs after fungal decay. X-ray diffraction analyses demonstrated that fungal decay increased the crystallinity of the poplar WFMRC but decreased the crystallinity of the larch WFMRC. Bore hole formation on the walls of wood cells, cell wall thinning, and collapsed cells were found in decayed samples, particularly in the poplar WFMRC. Although the WFMRCs could be classified as resistant to fungal decay, an appropriate protective measure should be considered to improve the outdoor durability of the WFMRC.
Article
The aim of this study was to investigate degraded birch trees (Betula pendula Roth) that suffered from a harmful fungus called Piptoporus betulinus. The main chemical analysis of B. pendula degraded by the fungus, included the holocellulose, alpha-cellulose, and lignin contents and was determined in cold and hot water and alcohol-benzene solubility in 1% NaOH mixtures. This fungus caused B. pendula to lose mass and chemical properties. The declining amount of holocellulose mass loss was 6.7% according to the holocellulose test. This decrement caused the quality of the birch holocellulose to decline. The total loss difference was 9.8% according to the alkaline solubility analysis of the 1% NaOH test and 14.3% according to the density analysis of the test. The loss difference was 4.2% according to the alcohol-benzene analysis of the test.
Article
Full-text available
Incipient decay of wood by brown-rot fungi causes measurable strength losses in wood before measurable weight loss occurs. Previous studies have shown that the high levels of strength loss that occur during incipient brown-rot decay may be related to loss in hemicellulose. This study investigates the effect of decay on hemicellulose composition and the relationship of decay to the mechanical properties of the wood. An in vitro test method was used to allow progressive sampling of southern pine exposed to monocultures of brown-rot fungi. The wood was subsequently analyzed by mechanical testing and chemical analysis. The results demonstrated a ratio of strength to weight loss of approximately 4:1. The chemical data indicated that early strength loss (up to 40%) was associated with loss of arabinan and galactan components. Subsequent strength loss (greater than 40%) was associated with the loss of the mannan and xylan components. Significant loss of glucan (representing cellulose) was only detected at greater than 75 percent modulus of rupture loss.
Article
Full-text available
Losses in toughness correlated with losses in weight were found in birch (Betula veralcosa), Scots pine (Pinus sylvestris) and Caribbean pine (Pinus caribaea var. hondurensis) over a six months period after inoculation with the blue stain fungus Lasiodiplodia theobromae. Differences between wood samples loaded radially and tangentially in the toughness testing device were not found, probably because the size of the samples (10x10x100mm) was smaller than that specified in standards. Higher strength losses at later stages of incubation were strongly correlated with degree of degradation of parenchyma cells producing weak planes of brash fracture and abrupt transwall failure. Transwall failure revealing the S-2 microfibrillar angle in early stages of incubation and a very smooth surface in more advanced stages of incubation, were also observed. Vessels in birch showed transwall failure after initial failure following degraded terminal parenchyma cells. Intrawall failure was also observed, particularly between the S-1-S-2 interface. Results indicate that this stain fungus could have a negative effect on the mechanical wood properties of the hard-and softwoods studied.
Article
Rapid depolymerization of cellulose occurs shortly after brown-rot fungi colonize wood. The chemical agent responsible for this initial depolymerization is most likely a low molecular weight compound (Cowling, 1961; Cowling and Brown, 1969) that diffuses through the crystalline microfibrils of cellulose, degrading the amorphous non-crystalline regions (Cowling and Brown, 1969; Highley, Palmer, Murmanis, 1983). It is important to identify the depolymerizing agent(s) produced by brown-rot fungi, because such information could serve as a foundation for the development of new methods to prevent wood decay.
Article
graphs, illus., tables
Article
Summary Sweetgum (Liquidambar styraciflua L.) wood blocks were decayed by Postia (= Poria) placenta in soil- block cultures. Decay was terminated at various weight losses, and the pore volumes available to four low molecular weight molecules, (water, 4 Å,; glucose, 8 Å,; maltose, 10 Å; and raffinose, 128,) and three dextrans (Mr 6,000, 38 Å; 11,200, 51 Å; and 17,500, 61 Å) were determined by the solute exclusion technique (Stone and Scallan 1968b). The volume in sound (undecayed) wood that was accessible to the seven probes varied from 1.0 ml g -1 for the three largest to 1.35 ml g -1 for water. Thus, the volume in sound wood attributable to lumens, pits, and other large openings was 1.0 ml g -1 and that accessible to water in the cell wall was 0.35 ml g -1 . Of this volume, 80% was inaccessible to molecules > 12 Å in diame- ter. As the wood was decayed, the volume of pores in the cell wall increased steadily to 0.7 ml g -1 at 35% weight loss. New cell wall volume was accessible to the four low molecular weight probes but not to molecules of Mr ≥ 6,000. The increase in accessible pore volume to the four smallest probes was gradual. Most of the new cell wall volume created by removal of components during decay was in the pore size range of 12 Å, to 38 Å. Within experimental error, no pores of > 38 Å, were observed in sound or decayed wood. Our results are consistent with the hypothesis that the initial depolymerization of cellulose, charac- teristic of brown rot, is caused by a diffusible agent. The molecular diameter of the agent is apparently in the range 12 Å, to 38 Å, and it causes erosion and thus enlargement of the pores to which it has access.
Article
It is not definitively known whether or not the production of extracellular hydrogen peroxide (H2O2) is a universal characteristic of brown-rot fungi. Cytochemical localization of H2O2 was tested in two brown-rot fungi, Tyromyces palustris and Coniophora puteana, by staining with cerium chloride, Transmission electron microscopy (TEM) showed the deposition of cerium perhydroxide within the fungal hyphae as well as wood cell walls affected by brown-rot fungi. TEM work indicated that extracellular H2O2 was present in brown-rot fungi and that H2O2 from brown-rot fungi diffused into the wood cell walls in the early stages of decay. The present work strongly suggests that H2O2 plays an important role in the early degradation of cellulose by brown-rot fungi. The usefulness of this technique for localizing H2O2 at high resolution with minimal nonspecific deposition is also discussed.