Sunteți pe pagina 1din 9

Polymer Degradation and Stability 89 (2005) 327e335

www.elsevier.com/locate/polydegstab

Composition, structure and thermal degradation


of hemp cellulose after chemical treatments
S. Ouajai, R.A. Shanks*
Applied Chemistry Department, RMIT University, GPO Box 2476V, Melbourne 3001, Australia
Received 23 November 2004; received in revised form 13 January 2005; accepted 18 January 2005
Available online 11 March 2005

Abstract

The thermal degradation behaviour of hemp (Cannabis sativa L.) fibres under a nitrogen atmosphere was investigated by using
thermogravimetry (TGA). The kinetic activation energy of treated fibres was calculated from TGA data by using a varied heating
rate from 2.5 to 30  C/min. The greater activation energy of treated hemp fibre compared with untreated fibre represented an
increase of purity and improvement in structural order. A hydrophobic solvent affected the degree of non-cellulosic removal.
Mercerisation and enzyme scouring removed non-cellulosic components from the fibre; however, structural disruption was observed
after higher alkaline concentration, 20 %wt/v and longer scouring time, respectively. Structural disruption was observed by X-ray
measurement. The FTIR results indicated an elimination of the non-cellulosic components by the mercerisation treatment and
a specific removal of low methoxy pectin by use of pectate lyase enzyme (EC 4.2.2.2). An increase of temperature at the maximum
rate of degradation and the rate of weight loss was characteristic of the purity and structure of treated hemp fibre.
Ó 2005 Elsevier Ltd. All rights reserved.

Keywords: Cellulose fibre; Thermal degradation; Kinetics; Scouring; Crystallinity

1. Introduction produce differences in non-cellulosic composition [3,4].


The hemp bundle bast fibres were found to contain a large
Composites derived from natural and sustainable amount of pectins (18%), hemicelluloses (16%) and
resources, especially cellulose are increasing in impor- a small amount of lignin (4%) [1,2]. These chemicals are
tance due to their numerous applications and advantages. not thermally stable and tend to degrade at an early stage
The composites require a strong fibre with good adhesion of heating. Further processing of a composite requires
between matrix and fibre to enhance their final properties. thermal stability information for materials selection and
The bast fibres from hemp (Cannabis sativa L.) were process operation. Removal of non-cellulosics from fibre
selected for pre-treatment. As a natural product, the surfaces was suggested to achieve this purpose. Various
complex fibre composite was created via biosynthesis. degrees of purity are required for different applications.
The bast fibres in hemp are bound by a central lamella and Therefore, several methods have been applied to hemp
arranged in bundles, separated by the cortex parenchyma cellulose.
cell with pectic- and hemicellulosic-rich cell wall [1,2]. Firstly, solvent extraction is an important method
The particular species, time of cultivation and weather conducted to remove the extractable fraction from
cellulosic fibres [3,4]. This procedure may cause slight
damage to the fibre structure and results in a more
* Corresponding author. Tel.: C61 3 9925 2122; fax: C61 3 9639
exposed cellulosic surface. Secondly, the chemical process
1321. of mercerisation is widely used to modify many types of
E-mail address: robert.shanks@rmit.edu.au (R.A. Shanks). cellulosic fibres. It is a well-known treatment for fibres

0141-3910/$ - see front matter Ó 2005 Elsevier Ltd. All rights reserved.
doi:10.1016/j.polymdegradstab.2005.01.016
328 S. Ouajai, R.A. Shanks / Polymer Degradation and Stability 89 (2005) 327e335

using alkaline solution, prior to composite formation. 2.2.2. Mercerisation


Most of the non-cellulosic components and part of the Dried fibres (2.5 g) were treated with various
amorphous cellulose can be removed by mercerisation. concentrations of NaOH solution (100 mL) and placed
The treatment not only changes the chemical composition in an oven at 30  C for 1 h in order to remove hemi-
of fibres but can rearrange or transform the crystalline celluloses and lignin associated with the fibres. The
structure of cellulose I to cellulose II, especially when concentrations of NaOH solutions were 3, 5, 8, 10, 12,
a high concentration of alkali has been applied. Thirdly, 15 and 20 %wt/v. The treatment with 8 %wt/v NaOH
a recent method is a biological treatment process. The was also conducted at 100  C. The alkaline treated
substrate of interest can be removed by a specific enzyme. fibres were subsequently washed with running tap water
Pectate lyase enzyme (EC 4.2.2.2) is recommended to followed by distilled water until no alkali was present in
remove low methoxy pectin. This process is an environ- the wash water.
mentally friendly process. Different treatments cause
a variation in the degree of impurities removed as well
as the degree of structural disruption. The effect of 2.2.3. Pectate lyase enzyme scouring
the difference in non-cellulosic composition and degree The untreated hemp fibres (5.0 g) were subjected to
of structural disruption on the thermal stability is an treatment with Scourzyme L (0.2, 0.5, 1.2, 2.5, 5.0,
important issue to be investigated. Several character- 10.0% on weight of fibre (owf)) in non-agitated 250 mL
isation techniques are suitable; however, the focus here is Erlenmeyer flask at 55  C for 0.5e24 h using a material
on thermogravimetry [5]. to liquor ratio of 1:3e1:100 and pH 8.5. Both enzyme
Thermogravimetry is one of the most widely used and substrate in citrateephosphate buffer solution were
techniques to monitor the composition [5] and structural preheated separately at 55  C for 10 min before mixing.
[6] dependence on the thermal degradation of natural In order to inactivate the enzyme, the reaction flasks
cellulose fibre. This is because the different compositions were chilled on ice for 10 min after the enzymatic
and supramolecular structures of cellulose behave treatment reached the desired time. The fibres were
differently when undergoing thermal degradation. The removed from solution and washed and then oven dried.
aim was to investigate the dependence of thermal
degradation on the applied purification methods;
2.2.4. Thermogravimetry
solvent extraction, mercerisation and enzyme scouring.
Dynamic experiments were performed using a Per-
Thermogravimetry was used to calculate the kinetic
kineElmer TGA7 instrument. Temperature programs
activation energy of cellulose degradation based on mass
for dynamic tests were from 35 to 850  C at a heating
loss of cellulose. The crystalline structure was observed
rate of 2.5e30  C/min. The measurements were con-
by X-ray scattering and the crystallinity index was
ducted under nitrogen (20 mL/min) and switched to air
calculated. In combination with X-ray scattering anal-
at 700  C.
ysis, an understanding of the structural and thermal
degradation relationship of cellulose fibres with treat-
ment methods was an objective. Other measurements 2.2.5. X-ray scattering analysis
such as FTIR and SEM were employed to assist with the The fibres (70 mg) were cut and pressed into a disk
interpretation of results. using a cylindrical steel mold (Ø Z 1.3 cm) with an
applied pressure of about 7000 kg/cm2 in a laboratory
2. Experimental press. Ni-filtered CuKa radiation (l Z 0.1542 nm) was
generated at 40 kV and 35 mA using a Bruker AXS D8
2.1. Materials WAXRDS. The X-ray diffractograms were recorded
from 5 to 60  2q (Bragg angle) by a goniometer
Hemp (Cannabis sativa L.) was obtained from equipped with scintillation counter at a scanning speed
Australian Hemp Resource and Manufacture (AHRM). of 0.02  /s and a sampling rate of 2 data/s.
The fibre obtained was a green-dried stalk after
decortication. Dry matter yield was 90% of field-dried
2.2.6. Fourier transform infrared (FTIR)
yields. Scourzyme L, pectate lyase (EC 4.2.2.2), was
spectroscopy measurements
kindly provided by Novozyme Australia Pty, Ltd.
The measurements were performed using a Perkine
2.2. Methods Elmer 2000 spectrometer. A total of 100 scans was
taken for each sample with a resolution of 2 cm1.
2.2.1. Solvent extraction The fibres were cut in an IKA MF10 cutting mill and
The fibres were subjected to Soxhlet extraction with sieved to provide a size range between 106 and 212 mm.
various solvents (acetone, benzene, ethanol and hexane) A mixture of 5.0 mg of dried fibres and 200 mg of KBr
for 3 h to remove any waxes present and then air-dried. was pressed into a disk for FTIR measurement.
S. Ouajai, R.A. Shanks / Polymer Degradation and Stability 89 (2005) 327e335 329

2.2.7. Morphological analysis a carboxylic acid rich fraction of pectin, because the
Scanning electron microscopy (SEM) was used to original vibration peaks of carboxylic acid and ester in
observe the microstructure and the surface morphology pectin were not resolved. In addition the acetyl group in
of treated and untreated hemp fibres. The instrument hemicelluloses occurred in same region [8]. Fortunately,
was a Phillips XL 30 Oxford 6650 SEM with an water-extracted pectin showed a characteristic of car-
acceleration voltage of 142 eV. The samples were coated boxylate ion [9]. The antisymmetric COO stretching
with gold to provide about 200 Å gold layer thickness was present at w1640 cm1. The carboxylate and ester
using a vacuum sputter coater. bands were well separated, leading to a measurable
content for each fraction. A gradual increase of 1640e
1733 cm1 absorbance ratio was obtained. This indi-
cates that after treatment the pectin was still present but
3. Results and discussion
with a higher degree of methyl ester content. This
demonstrates that the enzyme was specific for attack on,
3.1. FTIR results
and gradually removed the non-esterified fraction from
the structure of pectin. Furthermore, the higher content
The aim of using FTIR is to measure the change of
of methyl ester caused a reduction of the OH stretching
surface composition of fibres after treatment. Infrared
band (weaker H-bond). The band at 830 cm1 attrib-
spectra of hemp fibres after acetone extraction, mercer-
uted to an aromatic CeH out-of-plane vibration in the
isation and enzyme scouring are shown in Fig. 1. In
lignin was decreased in intensity after the acetone
general, the spectrum of the solvent treated hemp fibre is
extraction and mercerisation [10]. This indicated that
similar to that of the untreated hemp. However, the
the treatment reduced lignin content. This was contrary
vibration peak at 1733 cm1 attributed to the C]O
to the result exhibited by the enzyme treated fibres. The
stretching of methyl ester and carboxylic acid in pectin
other noticeable changes were an increase in intensity of
disappeared from mercerised fibres. This indicated the
the 897 cm1 bands attributed to the symmetric in-phase
removal of pectin and hemicelluloses by alkalisation.
ring-stretching mode and a decrease in intensity of the
Pectin contains both esterified and carboxylic acid
1431 cm1 band attributed to CH2 symmetric bending.
groups in the structure. Nevertheless, the FTIR spectra
The lateral crystallinity index of the fibres was evaluated
of enzyme treated hemp showed a band at 1733 cm1.
as the intensity ratio between IR absorptions at 1431
The presence of this band after treatment indicated the
and 897 cm1 assigned to the CH2 symmetric bending
existence of pectin [7]. The C]O band alone could not
mode and C1 group frequency, respectively [11]. The IR
reveal the difference in structure after removal of
lateral crystallinity index exhibited a variation with
treatment as shown in Table 1. The solvent extracted
and 8% NaOH treated fibres showed a slight decrease
of the index. After enzyme scouring, however, a slight
decrease of crystallinity index suggested that the
crystalline structure of the fibres was mildly disturbed
resulting in the presence of less ordered cellulose
structure.

3.2. X-ray scattering results

Fig. 2 shows X-ray diffractograms of the untreated,


acetone extracted, mercerised and bioscoured hemp

Table 1
The X-ray and IR crystallinity index of treated hemp fibres
Solvent Mercerisation Enzyme
extraction scouring
X-ray IR Conc. X-ray IR Time X-ray IR
(%wt/v) (h)
Raw 63.3 2.21 Raw 63.3 2.21 Raw 63.3 2.21
Hexane 66.4 1.13 3 75.5 1.85 0.5 72.0 1.50
Benzene 63.8 1.41 8 72.6 1.80 1.5 64.0 1.50
Acetone 67.4 2.15 12 66.4 0.63 6 57.2 1.33
Fig. 1. IR spectra of hemp fibres; (a) untreated, (b) acetone extracted,
Ethanol 60.4 1.41 20 58.2 0.88 24 58.0 1.35
(c) 8% NaOH treated and (d) 1.2% Scourzyme, 1.5 h treated.
330 S. Ouajai, R.A. Shanks / Polymer Degradation and Stability 89 (2005) 327e335

3.3. SEM

SEM images at magnifications between 500 and 600


were obtained for controlled and enzymatic treated
fibres reported in this work. Fig. 11(a) shows a fibre
bundle of untreated hemp covered by non-cellulosic
materials. The fibre bundle was 150e160 mm in di-
ameter. Fig. 11(b) indicated treatment with acetone
was not sufficient to remove all of the non-cellulosic
materials from the fibres. Only the solvent soluble
fractions were extracted. The fibre treated with NaOH
exhibited a considerably cleaner surface (Fig. 11(c)).
The surface of 1.2% Scourzyme L, 6 h treated fibres
(Fig. 11(d)) appears smoother with deeper inter-fibular
disintegration of the bundle. This indicated a sufficient
Fig. 2. X-ray diffractogram of treated hemp fibres. scouring time of 6 h. The disappearance of any non-
cellulosic components from the fibre seemed to have
a significant effect on the thermal stability. This will be
discussed in the following section.
fibres. The major diffraction planes of cellulose namely
110, 110, 012 and 020 are present at 14.8, 16.7, 20.7 and
22.5  2q angle [12]. Untreated fibre shows the character- 3.4. Thermal degradation of cellulose
istics of cellulose I. Solvent extraction by acetone caused
no change to the cellulose structure. However, the The differential thermogravimetry (DTG) curves
alkaline treatment with 8% NaOH caused an increase in (Fig. 3) of untreated hemp fibre show an initial peak
intensity of the 020 plane. As the concentration of between 50 and 160  C, which corresponds to a mass
NaOH reached 20 %wt/v, the crystalline transformation loss of absorbed moisture of approximately 5%. After
to cellulose II could be observed. The 20% NaOH this peak, the DTG curve showed three decomposition
treatment decreased the intensity of the 020 plane and steps: (1) the first decomposition shoulder peak at about
increased the intensity of the 1 10 and 012 planes. 250e320  C is attributed to thermal depolymerisation
A new 110 diffraction plane at the lowest 2q represented of hemicelluloses or pectin (mass loss 10%); (2) the
the introduction of cellulose II after treatment. This major second decomposition peak at about 390e400  C
cellulose structural change was likely to directly affect is attributed to cellulose decomposition (mass loss 55%);
the thermal degradation characteristic of the fibre. There (3) the small peak at 420  C (mass loss 30%) may be
was no crystalline transformation of the crystalline attributed to oxidative degradation of the charred
structure in the enzyme treated fibres. It is important to residue. The last peak in a nitrogen environment
note that the crystallinity index was used to indicate the occurred from the residue loss and occurred after
order of crystallinity rather than the crystallinity of switching gases from nitrogen to air. Decomposition in
crystalline regions [13]. This brought about the idea to
measure the changing of order of each crystalline plane
in cellulose separately. The crystalline order index was
determined from the fraction of the ratio of the 020
to the sum of 1 10, 012 and 020 reflection areas. A
deconvolution of the peak due to each diffraction plane
was achieved through curve fitting using a set of Pseudo-
Voigt curves to fit the experimental data. The X-ray
crystalline order index results are presented in Table 1.
There was a variation of results in the solvent extracted
sample. The crystallinity of fibres treated by 8% NaOH
was increased. The better packing or stress relaxation
was brought about by the removal of pectin as also
suggested by FTIR [14]. Higher concentration of NaOH
induced mercerisation of cellulose I into II, which
resulted in the decrease of crystallinity. After the enzyme
treatment, the results showed a reduction of the X-ray Fig. 3. The TG and DTG of untreated hemp fibre heated at 20  C/min
crystalline order index as a function of scouring time. in nitrogen and air.
S. Ouajai, R.A. Shanks / Polymer Degradation and Stability 89 (2005) 327e335 331

air was more complete and proceeded at a lower 3.4.2. Mercerisation


temperature than in nitrogen as a result the presence Results obtained for the hemp fibres after the
of free radicals and oxidation does include only free mercerisation treatment showed that mass loss de-
radicals [15]. Nevertheless, the reported TGA measure- pended on concentration of NaOH solution. The
ments were conducted under the inert atmosphere. mercerisation treatment lead to a fibre mass loss of
7.0, 9.6, 12.3, 14.1, 14.0, 14.7 and 15.6 %wt for NaOH
of 3, 5, 8, 10, 12, 15, and 20 %wt/v, respectively. The
3.4.1. Solvent extraction mass loss took into account the removal of soluble
The effect of solvent on the DTG peak position is matter during washing. The structure of cellulose
reported in Fig. 4. The extraction lowered the DTG transformed from cellulose I to cellulose II in 20%
maximum peak position compared with the untreated NaOH treated, but the 8% NaOH still maintained the
fibres. This decrease depended significantly on the original structure with a higher degree of crystallinity
hydrophobicity of the applied solvents. Benzene treated index. DTG results (Fig. 5) showed a thermal stability
fibres exhibited the lowest maximum decomposition change after pre-treatment. The main decomposition
temperature, followed by the hexane, acetone and temperature increased from 397.4  C (raw) to 410.3 and
ethanol treated fibres, respectively. Thus, the extracted 401.1  C for the 8% NaOH and 20% NaOH mercerised
composition that had the highest hydrophobicity fibres, respectively. The shoulder of the DTG peak at
property might be responsible for the thermal retarda- about 250e320  C disappeared after treatment, indicat-
tion of cellulose fibre degradation. Lignin as a compo- ing mass loss at this stage was mainly pectin and
nent of the fibres was degraded at a higher temperature. hemicelluloses. This corresponded well with the disap-
The structure of lignin is a highly aromatic polymer. pearance of the C]O band in the IR spectra of
Possibly, it was removed by benzene in larger amount mercerised fibres. According to the IR results, it was
than the other solvents. Moreover, no reduction of the apparent that this significantly affected the beginning of
shoulder peak was observed for all solvent extracted thermal degradation. The onset of degradation of
fibres. This indicated the removed compositions were mercerised fibre was improved compared with untreated
not of substances that degraded at a low temperature. fibre. This represented the removal of water-insoluble
Although solvent extraction can remove impurities from chemicals by alkaline reaction, which affected the main
the surface of the fibres; it could not improve the decomposition of cellulose. The decreased temperature
thermal stability of cellulose. No structure disruption of maximum degradation rate of the resulting fibres
was found in solvent extracted fibres. Solvent extracted indicated that a lower order of cellulose structure was
fibre had slight increase in crystallinity compared with present after strong alkaline solution treatment. This
the untreated fibres but it underwent thermal degrada- was confirmed by the reduction of X-ray crystallinity
tion at a lower temperature. This was probably due to index.
the loss of the lignin fraction that had a high thermal
stability. The occurrence of this material could prevent 3.4.3. Enzyme scouring
cellulose degradation and retain the degradation at TGA curves of water extracted pectin and treated
a high temperature [16]. fibre are presented in Fig. 6. The observed mass loss

Fig. 4. DTG of solvent extracted hemp fibres heated at 20  C/min in


nitrogen. Fig. 5. DTG of mercerised hemp fibre at 20  C/min in nitrogen.
332 S. Ouajai, R.A. Shanks / Polymer Degradation and Stability 89 (2005) 327e335

Fig. 6. DTG of bioscoured fibre heated at 20  C/min in nitrogen.


Fig. 7. DTG of various treated hemp fibres heated at 20  C/min in
nitrogen.

starting at 260  C was attributed to the decomposition


of light fractions (pectin and hemicelluloses) [16]. The lower temperature. In comparison with Scourzyme L
removal of pectin resulted in a significant reduction in treated fibres, the absence of mass loss belonging to low
mass loss at this temperature and a gradual shift of the methoxy pectin and a shift of the degradation temper-
maximum decomposition rate to higher temperature. ature to higher temperature were observed. Moreover,
This indicated a purity and thermal stability improve- the decomposition rate was increased. A further slight
ment with an increase in treatment time. According to increase of degradation temperature was observed in
the IR result for the extracted pectin, the type of pectin NaOH treated fibres. This was probably because more
present in the fibre was a pectate [9]. The thermal non-cellulosic material was removed and the high degree
stability enhancement may be attributed to the absence of structural order was retained. This revealed a re-
of low methoxy pectin, especially the one containing lationship between structure and the thermal degrada-
pectate salt [17]. The presence of an ion can lower tion of cellulose. A greater crystalline structure required
the degradation temperature. Nevertheless, the rate of a higher degradation temperature [18]. However, both
decomposition was increased. This could be explained non-cellulosic components and the crystalline order of
by a char created by the degradation of pectin. A slight cellulose played an important role in thermal degrada-
amount of pectin resulted in a low amount of char tion of the fibres.
formation. Generally, the presence of char could limit
the rate of decomposition of a material by delaying the
volatility rate of the gases produced. Hence, the removal 3.4.5. Kinetic measurement
of pectate caused a greater rate of cellulose mass loss. Fig. 8(a,b) represents the degradation of untreated
Although the structure of scoured fibre was slightly fibre at different heating rates. It has been suggested that
disrupted after treatment, the thermal stability was to avoid compensation effects in the estimation of the
improved. This indicated the presence of pectin had kinetic constants, different heating rates should be
a more significant effect than the slight structural considered [19]. Heating rates of 2.5, 5, 10, 15, 20,
disordering. However, this demonstrated the success of 30  C were chosen to study the thermal degradation
the enzyme scouring method that improved thermal kinetics of hemp fibre. A shift in the temperature of the
stability as well as preserving the structure of the maximum degradation rate occurred with increasing
cellulose. heating rate (Fig. 8(b)). The initial sample size of fibre
was controlled between 1.3 and 1.6 mg to avoid heat
3.4.4. Comparison of TGA after different treatments transfer problems at higher heating rates [20]. This
The comparison of thermal degradation between the observation can be seen from the constant mass loss
different treatments is shown in Fig. 7. The DTG of rates over the entire experimental range. From the
acetone treated fibres showed a reduction of tempera- kinetic evaluation, the major processes of degradation
ture at the maximum degradation rate of cellulose. The were considered, as indicated for the maximum temper-
compositions that degraded at low temperature were ature (DTG) in Fig. 8(b). The degradation of cellulose
still present after acetone extraction. This confirmed that was regarded as a first order-reaction [21]. Log(heating
the extracted composition was not the one degraded at rates) were plotted against 1/(temperature) for each
S. Ouajai, R.A. Shanks / Polymer Degradation and Stability 89 (2005) 327e335 333

specific conversion (Fig. 9). The slope of the obtained


linear plots was used to calculate a rate constant for the
thermal composition at each selected conversion. The
cellulose pyrolysis process cannot be described by single
activation energy over the whole pyrolysis range. Thus
in this work, the conversion from 15 to 65% was chosen.
Therefore, the activation energy over the main de-
composition region (300e400  C) was calculated using
an Arrhenius plot in accordance with Ozawa’s method
[22] and ASTM E1641-99 using the following equation:

EZ  ðR=bÞDðlog bÞ=Dð1=TÞ

where E is the activation energy, J/mol, b the approx-


imation derivative in 1/K, b the heating rates in K/min,
T the temperature (K) at constant temperature and R
the gas constant 8.314 J/(mol K). The activation energy
at various conversions is shown in Fig. 10. The obtained
activation energy (Ea) was in the range of 130e190 kJ/
mol. This depended upon the conversion and treatment
applied. Activation energy changed drastically from low
(15%) to high (45%) conversion and was then quite
stable at higher conversion. This indicated the degrada-
tion mechanism at low conversion was different from
that at high conversion. Possibly, the deviation at low
conversion indicated the cleavage of linkages with dif-
ferent bond energies. However, Scourzyme L treated
fibre showed the most stable activation energy over the
entire conversion. Untreated and mercerised fibres
showed a similar characteristic and trend. However,
they showed a large difference in activation energy at
low conversion with the enzyme treated fibre. The higher
Ea of enzyme treated fibre indicated greater purity than
Fig. 8. Raw hemp fibre heated at 2.5e30  C/min in nitrogen (a) TG,
(b) DTG. untreated fibre. Interestingly, after treatment by sodium
hydroxide solution, the degradation provided a low Ea
at low conversion similar to the untreated fibres,

Fig. 9. Arrhenius plots of logarithm of the heating rate versus the Fig. 10. Activation energy of raw and treated hemp fibres at various
reciprocal temperature at different percentage conversions. conversions.
334 S. Ouajai, R.A. Shanks / Polymer Degradation and Stability 89 (2005) 327e335

Fig. 11. SEM images of (a) raw, (b) acetone treated, (c) 8% NaOH treated and (d) 1.2% Scourzyme treated hemp fibres.

although most of the non-cellulosic components were this research. The FTIR results indicated a change of
removed and the fibre crystallinity was increased after non-cellulosic components in the treated fibres. The
treatment. The presence of an alkali ion can depress the X-ray crystallinity index depended on the method
thermal degradation [23]. The most probable function of applied and the treatment conditions. Thermogravim-
the alkali was to promote ionisation of hydroxyl groups etry revealed that thermal degradation of hemp
in the cellulose molecules [24]. This occurred at depended mainly on the cellulose structure and the
significantly lower conversion because the alkaline could content of non-cellulosic components that were present
access a lower ordered structure of cellulose. This part of in the fibre. The enzyme scoured fibres provided the
the structure degraded at lower temperature than the greatest improvement of purity and thermal stability, as
higher ordered part. Hence, it is possible that the indicated from SEM images and high degradation
presence of unremoved alkali metal ion produced this activation energy. The kinetic activation energy of
degradative characteristic of the fibres. Any remaining thermal degradation of the treated fibres varied with
metal ions may be present as salts with carboxylate conversion. Comparison between the methods for
groups of retained non-cellulosic carbohydrates or purification of natural fibres has shown that non-
lignin. cellulosic components are removed, depending on the
method employed, and the crystallinity and crystalline
form of the cellulose may be modified by the treatment
4. Conclusions or the absence of interaction from the extracted
component. The component of the fibres and the nature
The treatment of hemp fibre by solvent extraction, of the cellulose contribute significantly to the thermal
enzyme scouring and mercerisation was conducted in stability.
S. Ouajai, R.A. Shanks / Polymer Degradation and Stability 89 (2005) 327e335 335

Acknowledgements [10] Sun R, Lawther JM, Banks WB. Journal of Applied Polymer
Science 1996;62:1473e81.
[11] Nelson ML, O’Connor RT. Journal of Applied Polymer Science
The authors gratefully thank King Mongkut’s 1964;8:1325e41.
Institute of Technology North Bangkok (KMITNB), [12] Krassig H. Proceeding of the eighth cellulose conference. New
Thailand for a PhD scholarship. York: Syracuse; 1975. pp. 777e90.
[13] Mwaikambo LY, Ansell MP. Journal of Applied Polymer Science
2002;84:2222e34.
[14] Ray PK. Journal of Applied Polymer Science 1969;13:2593e600.
[15] Shafizadeh F, Bradbury AGW. Journal of Applied Polymer
References Science 1979;23:1431e42.
[16] Fisher T, Hajaligol M, Waymack B, Kellogg D. Journal of
[1] Vignon MR, Dupeyre D, Garcia-Jaldon C. Bioresource Tech- Analytical and Applied Pyrolysis 2002;62:331e49.
nology 1996;58:203e15. [17] Godeck R, Kunzek H, Kabbert R. European Food Research and
[2] Garcia-Jaldon C, Dupeyre D, Vignon MR. Biomass and Technology 2001;213:395e404.
Bioenergy 1998;14:251e60. [18] Yang P, Kokot S. Journal of Applied Polymer Science
[3] van der Werf HMG, Harsveld van der Veen JE, Bouma ATM, ten 1996;60:1137e46.
Cate M. Industrial Crops and Products 1994;2:219e27. [19] Meszaros M, Varhegyi G, Jakab E, Marosvolgy B. Energy Fuels
[4] Keller A, Leupin M, Mediavilla V, Wintermantel E. Industrial 2004;18:497e507.
Crops and Products 2001;13:35e48. [20] Volker S, Rickmann T. Journal of Analytical and Applied
[5] Muller-Hagedorn M, Bockhorn H, Krebs L, Muller U. Journal of Pyrolysis 2002;62:165e77.
Analytical and Applied Pyrolysis 2003;68e69:231e49. [21] Brandbury AGW, Sakai Y, Shafizadeh F. Journal of Applied
[6] Majdanac LD, Teodorovic MJ. Acta Polymerca 1987;38:661e6. Polymer Science 1979;23:3271e80.
[7] Bociek SM, Welti D. Carbohydrate Research 1975;42:217e26. [22] Ozawa T. Journal of Thermal Analysis 1970;2:301e24.
[8] Himmelsbach DS, Khalili S, Akin DE. Journal of the Science of [23] Tanczos I, Pokol G, Borsa J, Toth T, Schmidt H. Journal of
Food and Agriculture 2002;82:685e95. Analytical and Applied Pyrolysis 2003;68e69:173e85.
[9] Synytsya A, Copikova J, Matejka P, Machovic V. Carbohydrate [24] Michie RIC, Neale SM. Journal of Polymer Science: Part A
Polymers 2003;54:97e106. 1964;2:2063e83.

S-ar putea să vă placă și