You are on page 1of 15

Wood Science and Technology

https://doi.org/10.1007/s00226-018-1057-3

ORIGINAL

Effect of heat treatment on wood chemical composition,


extraction yield and quality of the extractives of some
wood species by the use of molybdenum catalysts

Marisabel Mecca1 · Maurizio D’Auria1 · Luigi Todaro2

Received: 24 January 2018


© Springer-Verlag GmbH Germany, part of Springer Nature 2018

Abstract
The effect of heat treatment and the presence of some molybdenum catalysts on
the amount of extractives in Populus nigra, Larix decidua, Paulownia tomentosa,
Castanea sativa and Quercus frainetto wood were studied. There are an increase
in the amount of lignin and extractives and a decrease in the amount of holocel-
lulose following thermo-treatment, in the treatment temperature range. Autoclave
treatment of wood with water in the presence of some molybdenum catalysts can
increase the amount of extracts, but it reduces solubility. Soxhlet extraction of wood
with ethanol/toluene mixture in the presence of ­H3PMo12O40 increases the amount
of extractives and their solubility in chloroform, while in the presence of ­MoO3 only
the solubility of extractives increases. GC–MS analysis of insoluble fraction showed
the presence of myo-inositol and some simple carbohydrates, mainly ribose, xylose
and glucose. GC–MS analysis of soluble fraction showed the presence of long-chain
acids and fatty acid esters of 10–20 carbon atoms, mainly decanoic acid, hexadeca-
noic acid and octadecanoic acid, which can be a source of fatty acids for biodiesel
production.

Electronic supplementary material  The online version of this article (https​://doi.org/10.1007/s0022​


6-018-1057-3) contains supplementary material, which is available to authorized users.

* Marisabel Mecca
marisabelmecca@libero.it
1
Department of Science, University of Basilicata, Viale dell’Ateneo Lucano 10, 85100 Potenza,
Italy
2
School of Agricultural Forestry, Food, and Environmental Science, University of Basilicata,
Viale dell’Ateneo Lucano 10, 85100 Potenza, Italy

13
Vol.:(0123456789)
Wood Science and Technology

Introduction

Heat treatment of wood is getting more and more popular over the last several
years, and it is one of the environmentally friendly methods to protect and to
extend the service life of wood. It is known that thermal modification enhances
overall dimensional stability and biological durability of wood along with chang-
ing its colour without using chemicals (Esteves and Pereira 2009). However, heat
treatment adversely influences mechanical properties of wood including compres-
sion strength, bending strength and hardness due to modification of the chemical
structure of its main components, namely hemicelluloses, cellulose, lignin and
extractives (Cademartori et al. 2013).
Although many studies have been carried out on the improvement of the over-
all properties of wood, there has been a recently increasing interest in the char-
acterization of the thermal degradation of treated wood products (Jones 2015;
Todaro et  al. 2015), which could be related to both increased demand for heat-
treated wood and newly developed environmental management policies (ISO
14040 2006; ISO 14044 2006). The thermal decomposition process of wood and
wood products is based on a series of complex reactions influenced by many fac-
tors, such as the heating rate, temperature, exposure time, pressure, moisture,
composition of biomass material and the size of particles.
Extractives are responsible for the colour, smell and durability of the wood
(Rowell et al. 2012). Furthermore, extractives can represent a source of valuable
fine chemicals such as fatty acids, terpenes and phenols. Therefore, the extrac-
tives can be used in chemical industry, in cosmetic industry or for their biological
antioxidant properties in pharmaceutical and nutraceutical industry.
Recently, several efforts have been reported relating to the lignin degradation
to obtain fine chemicals (Rinaldi et al. 2016).
While the use of oxidant catalysts in the lignin depolymerization is the object
of widespread research (Li et al. 2015; Mukherjee et al. 2016), the possible use
of these catalysts to induce oxidative depolymerization of wood, increasing the
amount of low molecular mass chemical compounds that can be obtained from
wood, has not been studied until now.
On the basis of these results, the possible use of some oxidants of Mo(VI) to
induce chemical degradation of wood was tested. The selection of this type of
oxidant species is due to the intention, on the one hand, to re-evaluate the work of
an ancient chemist, Francesco Mauro (chemist able to synthesize several Mo(V)
and Mo(VI) derivatives, born in Basilicata, and coworker of Stanislao Canizzaro)
(D’Auria et al. 2014), and on the other hand, on the basis of the green properties
of this type of catalyst (Podgors’ek et al. 2009).
Therefore, because extractives are the key constituents that provide the dura-
bility of wood, in this contribution, the objectives were to evaluate the effect of
heat treatment on extraction yield, the chemical composition and on the quality
of the extractives, to demonstrate that available new procedures could be used
to obtain wood degradation in order to (1) increase the amount of recoverable

13
Wood Science and Technology

extractives, and (2) show that wood can be a source of valuable fatty acids for the
production of biodiesel (Mecca et al. 2017).
For this purpose, Populus nigra L., Larix deciduas Mill., Paulownia tomen-
tosa Steud., Castanea sativa Mill. and Quercus frainetto Ten. untreated wood and
thermo-treated wood were used.
Populus nigra is a geographically widespread tree species of which stands are
found in large parts of Europe and Asia. The species is a typical pioneer species par-
ticularly in riverine areas. It is dioecious, heliophilous and characterized by an effi-
cient dissemination of both seeds and pollen, but also by its good ability to naturally
undergo vegetative propagation. It is a species of both social and economic interest.
In poplar breeding programs it plays a central role: 63% of the poplar cultivars from
the international register descend from this species, mainly as interspecific hybrids.
In addition to wood production, it is used as a windbreak for the protection of agri-
cultural land and for ornamental and landscaping purposes (Frison et al. 1995).
Larix decidua, a deciduous conifer endemic to Europe, is an industrially impor-
tant tree, characterized by a strong, water-resistant and durable wood but also flex-
ible in thin strips. The heartwood is particularly weatherproof and, therefore, mainly
used as construction timber. Due to its high resin content and difficult machinabil-
ity, its use for furniture production is limited (Pferschy-Wenzig et al. 2008). At pre-
sent, larch sawdust is mainly used for the production of pellet fuels, and it contains
phenolic compounds, such as lignans and flavonols (mainly dihydroquercetin and
dihydrokaempferol) (Salem et al. 2016). Barks and branches are waste material of
the timber industry with important, but also partially underexplored, supplies of bio-
logically active compounds; this material can be considered a good source of anti-
oxidants and phytoconstituents with possible use in cosmetic or nutraceutical prod-
ucts; furthermore, several conifer barks contain phytoconstituents, such as tannins,
terpenes and polyphenols (Bianchi et al. 2015).
Paulownia tomentosa is a plant species belonging to the Paulowniaceae family,
which includes 17 species of trees, all endemic to the South and East Asia. Paulow-
nia wood, a light-coloured hardwood, has been revered for centuries by Japanese
craftsmen because of its workability and beauty. In the Japanese tradition, Paulow-
nia was used to build kotos (Japanese harps) because of the wood’s superior acous-
tical quality, and it is now also used as timber and biomass, and in Chinese agro-
forestry systems because it grows fast, its wood is light but strong, its flowers are
rich in nectar, and its leaves make good fodder for farm animals. Paulownia wood
also resists splitting and deformation in the drying process and can retain nails and
screws without splitting (Kaygin et al. 2009). Fast developing programs for planting
and growing Paulownia, with the help of financial support, are performed in Italy.
Presently, they are the sources for the fast supply of high-quality timber and energy.
Castanea sativa, or sweet chestnut, is a plant species in the family Fagaceae,
native to Europe and Asia Minor. Castanea sativa heartwood is one of the most
durable wood species in Europe. Due to its good durability, the wood is frequently
used for utility poles and posts and other outdoor applications, both in and above
ground. In addition, this wood is used for manufacturing furniture, barrels and con-
structions, notably in southern Europe (Dieste et al. 2012). One of the reasons for
the good durability is the presence of extractives, predominantly tannins. Chestnut

13
Wood Science and Technology

trees are a rich source of tannins, which were used in the past for the production of
some medicines, leather production, as well as playing a part in the flavouring of
various types of alcoholic beverages (Gironi and Piemonte 2011).
Quercus is one of the most important woody genera due to its large number of
species, including trees and shrubs in Turkey. Oak belongs to the family Fagaceae,
its botanical name is Quercus frainetto, and it is known by many common names
across the globe such as Hungarian oak and Italian oak. These trees are also tolerant
to different environmental and climatic conditions and contribute to erosion control.
It is often used for firewood, and because of the rather high durability of its wood,
Quercus frainetto wood sometimes has been used as construction material in civil
engineering and mining. It is less suited for the manufacture of barrels and furniture
(Bartha et al. 1998).
Thermo-treatment of wood has been used in order to obtain a wood that is more
resistant to outdoor conditions, changing the decay resistance and the dimensional
stability; indeed, thermo-treated wood generally shows a decreased wettability.

Materials and methods

Sample preparation and heat treatment

Paulownia tomentosa, Populus nigra and Larix decidua boards free of defect for
each species were supplied by a local manufactures located at Calitri (Avellino prov-
ince, Campania Region, Italy).
Castanea sativa and Quercus frainetto boards were obtained from a natural high
forest located in Calvello municipality (Potenza province, Basilicata Region, Italy).
A total of eight sawn boards for each species (Populus nigra, Larix decidua,
Paulownia tomentosa, Castanea sativa and Quercus frainetto) with dimensions of
30 mm × 180 mm × 1500 mm (thickness × width × length), having a moisture content
of 20%, were used for the experiments.
The thermo-vacuum plant facility used for the test, following the description of
Ferrari et al. (2013b), was a semi-industrial prototype with an internal diameter of
1.7  m, modified to perform thermo-vacuum treatments at high temperature (up to
250 °C). The cylinder walls were heated by diathermic oil circulating between dou-
ble layers of steel.
The samples were dried to a moisture content of 0% in the same cylinder prior
to the thermo-vacuum treatments. The drying process of the samples having an ini-
tial moisture content of 20% was carried out at 100 °C and a pressure not exceed-
ing 25,000 Pa, corresponding to a water boiling temperature of 65 °C (Ferrari et al.
2013b).
In the next step, samples of the abovementioned species were exposed to the
thermo-vacuum treatment. Each thermo-vacuum treatment consisted of a heating
phase that started at a temperature of 100 °C and reached a maximum air tempera-
ture of 190 °C followed by a thermal treatment phase with a different constant air
temperature for 2 h: 170 °C for Castanea sativa, 200 °C for Populus nigra and Larix

13
Wood Science and Technology

decidua, 180 and 200 °C for Quercus frainetto, and 180 °C, 200 °C and 210 °C for
Paulownia tomentosa.
Three-hour cooling phase decreased the temperature of the samples to 100  °C.
The heating rate of the air was 12 °C/h. The temperature variation of the diathermic
oil, air and wood samples was measured using thermocouples. The reason for the
different treatment is that each type of wood is associated with one specific type of
treatment because of the different nature of hardwood and softwood (Ferrari et al.
2013a).

Soxhlet extraction

The samples of wood obtained in “Sample preparation and heat treatment” section
were dried at 105 °C overnight. Then, it was ground through a 40-mesh screen using
a Wiley mill.
The obtained material (1.0  g) was put into an extraction thimble, which was
placed into a Soxhlet extraction apparatus. The sample was extracted with 300 ml of
1:2 ethanol/toluene mixture (v/v) for 7 h. After this period, the solvent was evapo-
rated in vacuo (D’Auria et al. 2017).

Lignin content

The remaining sawdust was transferred to a 50-ml beaker, a cold H ­ 2SO4 solution
(72%) (15  ml) was added, and the mixture was frequently stirred for 2  h at room
temperature.
The mixture was then diluted to 3% (w/w) with 560 ml of distilled water, heated
under reflux for 4 h, filtered and washed with 500 ml of water. The residue was dried
at 105 °C to a constant mass. The holocellulose content was determined by differ-
ence between the residue amount after extraction and the lignin content (Mohareb
et al. 2012).

Autoclave extraction

A sample of wood (10 g) was put into an airtight glass jar with 50 ml of distilled
water and placed into the autoclave (Vapor Matic 770) for 20  min, at a tempera-
ture of 120  °C and a pressure of 1  bar for the extraction. The sample was filtered
and frozen at a temperature of − 28  °C. Then, it was lyophilized to remove water
(D’Auria et al. 2017). The obtained mixture was fractionated as follows: the mixture
was treated with chloroform (20  ml) and filtered, the solvent was evaporated, and
the residue was chromatographed using thin layer chromatography on silica gel in
the presence of a fluorescent indicator and using a hexane/ethyl acetate mixture. The
different bands revealed by UV irradiation were scraped off and eluted with ethyl
acetate. The solvent was evaporated, and the residue was analysed as described in
“Gas chromatographic–mass spectrometric analyses” section. The chloroform-insol-
uble fraction was treated as described in “Derivatization” section.

13
Wood Science and Technology

Synthesis of ­MoO3 supported on silica catalyst

The calculated amount of 5 g of aqueous ammonium heptamolybdate solution [81


wt% as ­MoO3 (4.05 g)] was added into a suspension of 25 g of fumed silica (Aerosil
380) in distilled water (16% M­ oO3/fumed silica). The solution was then evaporated
to dryness at 100 °C in a rotary evaporator. The solid obtained was dried at 120 °C
overnight and calcinated at 650 °C for 5 h in a muffle furnace, getting about 25 g of
­MoO3/SiO2 (Suzuki et al. 1994). XPS analysis of the catalyst showed a Mo 3d5/2/Si
2p ratio of 2.97, in agreement with the value of 3.06 reported in the literature. The
signal of Mo 3d was found at 233.3 eV, while that of Si 2p was detected at 103.8 eV.
XPS spectra were acquired with a LH X1 Leybold (Germany) instrument using a
dual achromatic Mg Kα(1,2) (1253.6 eV) and a Al Kα(1,2) (1486.6 eV) source operat-
ing at a constant power of 260 W. Wide and detailed spectra were collected using
the fixed analyser transmission mode of operation with a channel width of 1.0 or
0.1 eV, respectively, and pass energy of 50 eV. Under these conditions, the instru-
mental contribution to the line width was kept constant. The measured full width at
half maximums of the Au ­4f7/2 (84.0 eV) and Cu 2p3/2 (932.7 eV) signals used for
calibration purposes were 1.3 and 1.6 eV, respectively.

Synthesis of ­H3PMo12O40 catalyst

The heteropolyoxometalate ­H3PMo12O40 was prepared according to the literature


procedure (Tayebee and Alizadeh 2007). 3.4 ml of H ­ 3PO4 (85% m/m) and 142 ml of
­HClO2 12 M were added successively to 210 ml solution of ­Na2MoO4 2.85 M. The
warm solution took on a yellow colour, and the disodium salt ­Na2H[PMo12O40] pre-
cipitated. After cooling to room temperature, the microcrystalline powder was fil-
tered and air-dried. The salt was then recrystallized from a (20 ml/100 ml) E
­ t2O/H2O
mixture obtaining about 50 g of a greenish precipitate. ­H3PMo12O40 was obtained
from a solution of 50  g of ­Na2HPMo12O40, just recrystallized, in 100  ml of ­H2O,
acidified by 25  ml of HCl 12  M and extracted with 150  ml of E ­ t2O. When water
was added, yellow crystals of ­H3PMo12O40 precipitated. The solid so obtained was
filtered and dried at 70–80 °C (Rahimi et al. 2014). FT-IR spectrum of the catalyst
showed peaks at νmax 750, 830, 952 and 1064  cm−1, in agreement with reported
data (Javidi et  al. 2014). Infrared spectra were obtained utilizing a Bruker Alpha
FT-IR spectrophotometer (Bruker Photonics, Billerica, MA) configured for attenu-
ated total reflectance at ambient temperature.

Autoclave and Soxhlet extraction in the presence of Mo catalysts

The sample (10  g) was put into an airtight glass jar with 50  ml of distilled water
and 1 g of the catalysts ­MoO3/SiO2 or ­H3PMo12O40, and then it was placed into the
autoclave (Vapor Matic 770) for 20 min at a temperature of 120 °C and a pressure
of 1 bar. Then, the sample was filtered, frozen at a temperature of − 28 °C and lyo-
philized to remove water, while for Soxhlet extraction, the sample (1.0 g) was put
into an extraction thimble with 0.1  g of the catalysts M
­ oO3/SiO2 or H­ 3PMo12O40,

13
Wood Science and Technology

and then it was placed into a Soxhlet extraction apparatus. The sample was extracted
with 300 ml of 1:2 ethanol/toluene mixture (v/v) for 7 h. After this period, the sol-
vent was evaporated in vacuo.

Derivatization

To about 100  mg of the chloroform-insoluble fraction of the extractives, 1  ml of


pyridine and 1 ml of acetic anhydride were added, and the sample was allowed to sit
at room temperature for 48 h. Then, the solvent was exchanged with ethanol under
reduced pressure followed by drying in vacuo (D’Auria et al. 2017). The residue was
chromatographed using thin layer chromatography on silica gel in the presence of a
fluorescent indicator. The different zones revealed by UV irradiation were separated
and eluted with ethyl acetate. The solvent was evaporated, and the residue was ana-
lysed as described in “Gas chromatographic-mass spectrometric analyses” section.

Gas chromatographic–mass spectrometric analyses

Analyses of all the extractives obtained using all the procedures previously described
were accomplished with an HP 6890 Plus gas chromatograph equipped with a Phe-
nomenex Zebron ZB-5 MS capillary column (30  m × 0.25  mm i.d. × 0.25  μm FT)
(Agilent, Milan, Italy). An HP 5973 mass selective detector (Agilent) was utilized
with helium at 0.8  ml/min as the carrier gas. A split injector was maintained at
250 °C and the detector at 230 °C. The oven was held at 80 °C for 2 min, then grad-
ually warmed, 8 °C/min, up to 250 °C and held for 10 min. Tentative identification
of aroma components was based on mass spectra and Wiley 6 and NITS 98 library
comparison.

Results and discussion

Populus nigra, Larix decidua, Paulownia tomentosa, Castanea sativa and Quercus
frainetto samples were used in this study. Samples of these woods were thermo-
treated as described in “Materials and methods” section, and the amount of extrac-
tives, lignin and holocellulose was determined.
Heat treatment of wood changes its chemical composition by degrading cell wall
compounds and extractives. The chemical changes due to heating depend on the
duration and temperature of the treatment, the temperature being the main factor
(Bourgois et al. 1989). For low temperatures, the wood dries, beginning with the loss
of free water and finishing with bound water, whereas at high temperatures the car-
bonization processes start with the formation of ­CO2 and other pyrolysis products.
The histograms in Fig. 1 (where the controls are shown in white and the thermo-
treated samples in black) show that the holocellulose content of the treated samples
decreased compared to those of the untreated wood. In contrast, the lignin content
of the samples is relatively higher in the treated samples than in the untreated sam-
ples. This trend was observed for all samples, except for Populus nigra heat-treated

13
Wood Science and Technology

(a) % HOLOCELLULOSE
Control
Treated
64.8
61.6 60.8
56.7
53.7 54.5
51.1 51.9
50.3
45.5 45.7
42.1
36.2

% LIGNIN 54.5
(b)
48.1
45.1 45.2

39.6 40.4
39.0 39.5
37.7 37.6

32.7 31.9
27.8

Fig. 1  Content of holocellulose (a) and lignin (b) of the samples

at 200  °C and Paulownia tomentosa thermally treated at 180  °C, where an oppo-
site pattern was observed: the content of holocellulose increased and the content of
lignin decreased.

13
Wood Science and Technology

The hemicelluloses are the first structural compounds to be thermally affected,


even at low temperatures. The degradation starts by deacetylation, and the
released acetic acid acts as a depolymerization catalyst that further increases pol-
ysaccharide decomposition (Tjeerdsma et al. 1998). The content of carbohydrates
decreases with the severity of the treatment and depends on the wood species.
Esteves et al. (2008) determined the content of sugars by acid hydrolysis before
and after heat treatment and concluded that hemicelluloses were affected first, as
manifested by diminishing yields of xylose, arabinose, galactose and mannose. At
higher temperatures, indeed, xylose and mannose content in wood decreases, and
arabinose and galactose disappear (Jamsa and Viitaniemi 2001).
Cellulose is less affected by the heat treatments, probably because of its crys-
talline nature.
The change in lignin content is recognized to be strongly dependent on the
temperature, treatment time, and on the characteristics of the wood species (Hill
2007). It is generally accepted that the increase in lignin content in the thermo-
treated wood is due to the loss of polysaccharide material that occurs during heat-
ing. However, as suggested by Esteves and Pereira (2009), the increase in lignin
content of the heat-treated wood cannot be considered pure lignin since polycon-
densation reactions occur in the cell wall of other wood components, resulting
in polymerization and in further cross-linking, thus increasing the apparent con-
tent of lignin. The increase in cellulose content of Populus nigra and Paulownia
tomentosa after the thermal treatment is a new behaviour that can be explained
only by assuming that in those species, the thermo-treatment induced a prefer-
ential lignin degradation with consequent decrease also in compounds deriving
from lignin, such as the extractives (see below).
The extractives were determined by using a Soxhlet procedure in ethanol/tolu-
ene or by extraction with water in an autoclave (D’Auria et al. 2017).
The cleavage of the ether linkages, especially β-O-4, leads to the formation
of free phenolic hydroxyl groups and α- and β-carbonyl groups (Nuopponen
et al. 2004), which are responsible for cross-linking via formation of methylenic
bridges. The methoxyl content decreases, and the new reactive sites on the aro-
matic ring can lead to further condensation reactions (Wikberg and Maunu 2004).
Most of the extractives disappear or degrade during the heat treatment, espe-
cially the most volatile ones.
The extraction of the samples in the autoclave resulted in a reduced amount
of extractives with respect to Soxhlet extraction in ethanol/toluene, as shown in
Fig. 2, demonstrating that the content of the extractives also depends on the dif-
ferent solvents used.
Following the heat treatment, the content of extractives (ethanol/toluene)
increased by 8.26% for Castanea s., by 5.66% and 4.80% for Quercus f. at 180 °C
and 200 °C, respectively, by 0.49% for Larix d. and by 7.44% and 2.34% for Pau-
lownia t. at 200 °C and 210 °C, respectively. On the contrary, after the extraction
by using Soxhlet apparatus, the amount of the extractives decreased by 0.60%
for Paulownia t. at 180  °C and by 2.96% for Populus n., compared to those of
untreated wood.

13
Wood Science and Technology

% EXTRACTIVES
(a) Control
Treated

15.7
15.3

10.2
9.4
8.5 8.6
7.9
7.4 7.3

5.7

3.7 3.4
2.9

% EXTRACTIVES
(b) 7.6

5.3
4.8
4.2

3.3

1.6 1.6
1.4 1.4
0.9 1.0 0.9
0.5

Fig. 2  Amount of extractives by Soxhlet (a) and autoclave (b) extraction of the samples

When the extraction was performed in an autoclave using water, the content
of the extractives increased by 3.40% for Castanea s., by 0.20% for Larix d., by
1.10% for Populus n., and by 0.14%, 0.59% and 0.07% for Paulownia t. at 180 °C,

13
Wood Science and Technology

200  °C and 210  °C, respectively, but the amount of the extractives decreased by
0.50% and 2.00% for Quercus f. at 180  °C and 200  °C, respectively, compared to
those of untreated wood. As reported below, the composition of water extractives
mainly depends on the presence of pentoses and hexoses. It is reasonable that in
thermo-treated species where the cellulose amount increases, the fraction that can be
extracted in water increases.
In addition, the possible effect of oxidants of Mo(VI) to induce degradation of
wood during the autoclave and Soxhlet extraction was tested. The effect of the pres-
ence of 10% of a polyoxometalate compound, such as ­H3PMo12O40 and that of silica-
supported ­MoO3, was tested in particular for Castanea sativa at 170 °C. The results
are reported in Fig.  3. Clearly, the presence of the Mo(VI) derivatives induced an
extensive degradation of wood, thereby allowing an increased recovery of extrac-
tives. In particular, in Soxhlet H
­ 3PMo12O40 increased the amount of extractives by
3.4% and by 8.2% in the autoclave. ­MoO3 reduced the amount of extractives by 0.9%
in Soxhlet, but increased the extraction yield by 1.2% in the autoclave. Furthermore,
the obtained extractives were fractionated considering the capability of the com-
ponents to be soluble in chloroform (dielectric constant 4.81, able to dissolve apo-
lar and medium polar compounds). In the case of the present samples, in Soxhlet
extraction in the presence of H­ 3PMo14O40 the solubility of the extracts increased by
9.2%, while it was reduced by 1.5% when the extraction was performed in the auto-
clave. These results can be explained by considering the different effects of the oxi-
dant on lignin and cellulose. The polyoxometalated compound induced a selective
degradation of lignin and cellulose inducing an increase in the amount of extractives

% EXTRACTIVES Castanea S. 170°C


Soxhlet
Chloroform Solubility
19.1
Autoclave

15.7 15.5 15.8


14.8

12.4

8.8
7.6

3.2
2.7
1.2
0.3
Control H3PMo12O40 MoO3 Control H3PMo12O40 MoO3

Fig. 3  Amount of extractives and solubility of Castanea sativa at 170 °C by Soxhlet (white) and auto-
clave (grey) extraction in the presence of the Mo polyoxometalate derivative

13
Wood Science and Technology

both in Soxhlet and in autoclave extraction processes. However, the autoclave treat-
ment allows the selective extraction of polar components of the extracts, with the
consequent reduction in the chloroform-soluble fraction. When the Soxhlet extrac-
tion was performed in the presence of ­MoO3, the oxidant was not able to oxidize
wood, and the treatment induced only an oxidation process of the extractives that
are reduced. The oxidation occurred mainly on the polar fraction of the extractives,
and then, the solubility in chloroform increased by 12.3%. When the extraction was
performed in the autoclave, ­MoO3 was able to induce an oxidative depolymerization
of wood with the consequent increase in the extractives amount. However, it reduces
the solubility in chloroform by 2.4% considering that, in water, the extraction occurs
mainly on polar compounds.
The main compounds of the chloroform-soluble fraction of the extractives were
determined on the basis of their mass spectra. These compounds were obtained with
an overall chemical yield of 31–44% (w/w), based on the different species of wood.
From the GC–MS analysis, the main compounds of the soluble extractives are long-
chain acids and fatty acids or esters of fatty acids from 10 to 20 carbon atoms, spe-
cially n-hexadecanoic acid, 9-octadecenoic acid and 9,12-octadecadienoic acid. On
the contrary, using wood of conifers, tall oil obtained in the kraft pulp production
contains large amounts of fatty acids (Rowe 1989) (see Supplementary Material).
The main compounds of the insoluble fraction of extractives were determined
after acetylation of the mixture. Sugars, in particular glucopyranose pentaacetate
and glucitol hexaacetate, were recovered. Furthermore, a polyhydroxylated com-
pound, myo-inositol hexaacetate, was determined in the mixture. The overall chem-
ical yields were in the range of 11–63% (w/w), based on the different species of
wood.
There also appeared new compounds in the thermo-treated samples resulting
from the degradation of cell wall structural components: apocynin (a, S1) in Cas-
tanea sativa at 170 °C and Populus nigra at 200 °C, with anti-arthritic (Hart et al.
1990) and anti-asthmatic (Stefanska and Pawliczak 2008) properties, aspidinol (b,
S2) in Quercus frainetto at 200  °C, that is an anti-parasite (Bosman et  al. 2004),
sinapaldehyde (c, S3) in Quercus frainetto at 180 °C and Populus nigra at 200 °C,
used as a flavouring agent, and it also is a cyclooxygenase inhibitor (Setzer 2011),
and fenozan (d, S4) and asarinin (e, S5) in Paulownia tomentosa at 180 °C, 200 °C
and 210  °C, which are both antioxidants, vasodilators and antithrombotic agents
(Bespalov et al. 2016; Yasuda et al. 2012; Mordi et al. 2016) (Fig. 4 and S1, S2, S3,
S4, S5 in Supplementary Material).

Conclusion

In the experimental conditions, heat treatment changed the chemical composition of


wood by degrading cell wall compounds and extractives. Hemicelluloses were the
most affected compounds, whereas cellulose was more resistant to heat.
As a result, heat treatment increased the yield of extractives and lignin and
reduced the amount of holocellulose. This trend was observed for all sam-
ples, except for Populus nigra heat-treated at 200  °C and Paulownia tomentosa

13
Wood Science and Technology

(a) (b) (c)

(d) (e)

Fig. 4  New compounds of thermo-treated sample

thermally treated at 180 °C, where an opposite pattern was observed: the content
of holocellulose increased, and the content of lignin decreased.
In this research, the possible effect of oxidants of Mo(VI) to induce degrada-
tion of wood during the autoclave and Soxhlet extraction was tested particularly
for Castanea sativa wood thermo-treated at 170 °C. As a general statement, the
polyoxometalated compound induced a selective degradation of lignin and cellu-
lose. It increased the extractives amount both in Soxhlet and in autoclave extrac-
tion processes, with a greater activity on heat-treated wood.
About autoclave and Soxhlet treatment of the samples in the presence of some
molybdenum catalysts, in Soxhlet extraction both catalysts increased solubility,
but only H­ 3PMo12O40 increased the extraction yields, while in the autoclave both
catalysts increased the yields of extractives, but they reduced solubility. Clearly,
in these conditions, the presence of the oxidants induced a selective degradation
of cellulose increasing the amount of chloroform-insoluble carbohydrates.
The effect of the oxidants seems to induce mainly the oxidative fission of cel-
lular membranes. The use of a molybdenum catalyst depended on the evidences
that molybdenum can act as catalyst in Fenton-like reactions. However, molybde-
num compounds have not been used for this purpose until now.
The organic solvent-soluble fraction was composed of 31–44% by long-chain
acid and fatty acid esters, while the water-soluble fraction of the extractives and
the insoluble fraction of extractives can be a source of simple carbohydrates.

13
Wood Science and Technology

In conclusion, the heat treatment and the use of molybdenum catalysts increased
fatty acid esters which have a possible use in biodiesel synthesis. Furthermore, the
increased amount of carbohydrates allowed to obtaining valuable amounts of bio-
logically active compounds.

References
Bartha D, Roloff A, Weisgerber H, Lang UM, Stimm B, Schütt P (1998) Enzyklopädie der Holzgewächse:
Handbuch und Atlas der Dendrologie (Encyclopedia of woody plants: handbook and atlas of den-
drology). Wiley-VCH, Weinheim
Bespalov VG, Alexandrov VA, Korman DB, Baranenko DA (2016) Phenozan, a synthetic phenolic anti-
oxidant, inhibits the development of spontaneous tumors in rats and mice. Drug Res 66(09):489–494
Bianchi S, Kroslakova I, Janzon R, Mayer I, Saake B, Pichelin F (2015) Characterization of condensed
tannins and carbohydrates in hot water bark extracts of European softwood species. Phytochemistry
120:53–61
Bosman AA, Combrinck S, Roux-Van der Merwe R, Botha BM, McCrindle RI, Houghton PJ (2004) Iso-
lation of an anthelmintic compound from Leucosidea sericea. S Afr J Bot 70(4):509–511
Bourgois J, Bartholin MC, Guyonnet R (1989) Thermal treatment of wood: analysis of the obtained prod-
uct. Wood Sci Technol 23:303–310
Cademartori PHG, dos Santos PS, Serrano L, Labidi J, Gatto DA (2013) Effect of thermal treatment on
physicochemical properties of Gympie messmate wood. Ind Crops Prod 45:360–366
D’Auria M, Colella C, Masini N (2014) Francesco Mauro-un chimico lucano (Francesco Mauro-a lucan
chemist). Edizioni Scientifiche Italiane, Napoli
D’Auria M, Mecca M, Todaro L (2017) Chemical characterization of cedrus deodara wood extracts using
water and molybdenum catalysts. J Wood Chem Technol 37:163–170
Dieste A, Rodrıguez K, Bano V (2012) Wood–water relations of chestnut wood used for structural pur-
poses. Eur J Wood Prod 71:133–134
Esteves B, Pereira H (2009) Wood modification by heat treatment: a review. Bioresources 4:370–404
Esteves B, Graça J, Pereira H (2008) Extractive composition and summative chemical analysis of ther-
mally treated eucalypt wood. Holzforschung 62(3):344–351
Ferrari S, Allegretti O, Cuccui I, Moretti N, Marra M, Todaro L (2013a) A revaluation of turkey oak
wood (Quercus cerris L.) through combined steaming and thermo-vacuum treatments. Bioresources
8:5051–5066
Ferrari S, Cuccui I, Allegretti O (2013b) Thermo-vacuum modification of some European softwood and
hardwood species treated at different conditions. Bioresources 8:1100–1109
Frison E, Lefèvre F, De Vries S, Turok J (1995). Populus nigra network. Report of the first meeting, 3–5
Oct 1994, Izmit, Turkey. IPGRI, Rome, Italy
Gironi F, Piemonte V (2011) Temperature and solvent effects on polyphenol extraction process from
chestnut tree wood. Chem Eng Res Des 89:857–862
Hart BA, Simons JM, Shoshan KS, Bakker NP, Labadie RP (1990) Antiarthritic activity of the newly
developed neutrophil oxidative burst antagonist apocynin. Free Radic Biol Med 9(2):127–131
Hill C (2007) Wood modification: chemical, thermal and other processes. Wiley, Chichester
ISO 14040 (2006) Environmental management: life cycle assessment—principles and framework. Inter-
national Organization for Standardization, Geneva
ISO 14044 (2006) Environmental management: life cycle assessment—requirements and guidelines.
International Organization for Standardization, Geneva
Jamsa S, Viitaniemi P (2001) Heat treatment of wood better durability without chemicals. In: Rapp AO
(ed) Review on heat treatments of wood. Cost Action E22. Proceedings of the special seminar,
Antibes, France, pp 17–22
Javidi J, Esmaeilpour M, Rahiminezhad Z, Dodeji FN (2014) Synthesis and characterization of
­H3PW12O40 and ­H3PMo12O40 nanoparticles by a simple method. J Cluster Sci 25:1511–1524
Jones JC (2015) Letter: fire loads and the calorific value of wood. Fuel 159:975
Kaygın B, Gündüz G, Aydemir D (2009) Some physical properties of heat-treated paulownia (Paulownia
elongata) wood. Dry Technol 27:89–93

13
Wood Science and Technology

Li C, Zhao X, Wang A, Huber GW, Zhang T (2015) Catalytic transformation of lignin for the production
of chemicals and fuels. Chem Rev 115:11559–11624
Mecca M, Todaro L, D’Auria M (2017) Extractives from cedar deodara and alnus cordata in the pres-
ence of molybdenum catalysts. Chem Sel 2:2536–2538
Mohareb A, Sirmah P, Petrissans M, Gerardin P (2012) Effect of heat treatment intensity on wood chemi-
cal composition and decay durability of Pinus patula. Eur J Wood Prod 70:519–524
Mordi RC, Fadiaro AE, Owoeye TF, Olanrewaju IO, Uzoamaka GC, Olorunshola SJ (2016) Identifica-
tion by GC–MS of the components of oils of banana peels extract, phytochemical and antimicrobial
analyses. Res J Phytochem 10(1):39–44
Mukherjee A, Mandal T, Ganguly A, Chatter PK (2016) Lignin degradation in the production of bioetha-
nol: a review. ChemBioEng Rev 3:86–96
Nuopponen M, Vuorinen T, Jamsä S, Viitaniemi P (2004) Thermal modifications in softwood studied by
FT-IR and UV resonance Raman spectroscopies. J Wood Chem Technol 24(1):13–26
Pferschy-Wenzig EM, Kunert O, Presser A, Bauer R (2008) In vitro anti-inflammatory activity of larch
(Larix decidua L.) sawdust. J Agric Food Chem 56:11688–11693
Podgorsˇek A, Zupan M, Iskra J (2009) Oxidative halogenation with “green” oxidants: oxygen and hydro-
gen peroxide. Angew Chem Int Ed 48:8424–8450
Rahimi R, Fatemeh R, Mahboubeh R (2014) Synthesis characterization and photocatalytic activity of
porphyrin-polyoxometalate hybrid material. The 18th international electronic conference on syn-
thetic organic chemistry. Multidisciplinary Digital Publishing Institute, Basel, pp 3–4
Rinaldi R, Jastrzebski R, Clough MT, Ralph J, Kennema M, Bruijnincx PCA, Weckhuysen BM (2016)
Paving the way for lignin valorisation: recent advances in bioengineering, biorefining and catalysis.
Angew Chem Inter Ed 55:8164–8215
Rowe JW (ed) (1989) Natural product of woody plants. Springer, Berlin, pp 952–978
Rowell RM, Pettersen R, Han JS, Rowell JS, Tshabalala MA (2012) Cell wall chemistry. In: Rowell RM
(ed) Handbook of wood chemistry and wood composites. CRC Press, Boca Raton, pp 33–72
Salem MZM, Elansary HO, Elkelish AA, Zeidler A, Ali HM, Hefny MEL, Yessoufou K (2016) In vitro
bioactivity and antimicrobial activity of picea abies and Larix decidua wood and bark extracts.
Bioresources 11:9421–9437
Setzer WN (2011) Lignin-derived oak phenolics: a theoretical examination of additional potential health
benefits of red wine. J Mol Model 17(8):1841–1845
Stefanska J, Pawliczak R (2008) Apocynin: molecular aptitudes. Mediat Inflamm 2008:106507
Suzuki K, Hayakawa T, Shimizu M, Takehira K (1994) Partial oxidation of methane over silica supported
molybdenum oxide catalysts. Cat Lett 30:159–169
Tayebee R, Alizadeh MH (2007) Water as an efficient solvent for oxygenation transformations with 34%
hydrogen peroxide catalyzed by some heteropolyoxometalates. Monatsh Chem 138:763–769
Tjeerdsma BF, Boonstra M, Pizzi A, Tekely H, Millitz H (1998) Characterisation of thermally modified
wood: molecular reasons for wood performance improvement. Holz Roh Werkst 56:149–153
Todaro L, Rita A, Cetera P, D’Auria M (2015) Thermal treatment modifies the calorific value and ash
content in some wood species. Fuel 140:1–3
Wikberg H, Maunu SL (2004) Characterisation of thermally modified hard- and softwoods by 13C
CPMAS NMR. Carbohydr Polym 58(4):461–466
Yasuda K, Ikushiro S, Wakayama S, Itoh T, Yamamoto K, Kamakura M, Munetsuna E, Ohta M, Sakaki
T (2012) Comparison of metabolism of sesamin and episesamin by drug-metabolizing enzymes in
human liver. Drug Metab Dispos 40(10):1917–1926

Publisher’s Note  Springer Nature remains neutral with regard to jurisdictional claims in published
maps and institutional affiliations.

13

You might also like