You are on page 1of 9

108 Die Angewandte Makromolekulare Chemie 272 (1999) 108–116 (Nr.

4753)

The effect of chemical treatment on the properties of hemp, sisal,


jute and kapok for composite reinforcement

L. Y. Mwaikambo*, M. P. Ansell
Department of Materials Science and Engineering, University of Bath, Bath BA2 7AY, UK

SUMMARY: Two chemical treatments were applied to hemp, sisal, jute and kapok natural fibres to create
better fibre to resin bonding in natural composite materials. The natural fibres have been treated with varying
concentrations of caustic soda with the objective of removing surface impurities and developing fine struc-
ture modifications in the process of alkalisation. The same fibres were also acetylated with and without an
acid catalyst to graft acetyl groups onto the cellulose structure, in order to reduce the hydrophilic tendency of
the fibres and enhance weather resistance.
Four characterisation techniques, namely XRD, DSC, FT-IR and SEM, were used to elucidate the effect of
the chemical treatment on the fibres. After treatment the surface topography of hemp, sisal and jute fibres is
clean and rough. The surface of kapok fibres is apparently not affected by the chemical treatments. X-ray
diffraction shows a slight initial improvement in the crystallinity index of the fibres at low sodium hydroxide
concentration. However, high caustic soda concentrations lower the fibre crystallinity index. Thermal analy-
sis of the fibres also indicates reductions in crystallinity index with increased caustic soda concentrations and
that grafting of the acetyl groups is optimised at elevated temperatures. Alkalisation and acetylation have
successfully modified the structure of natural fibres and these modifications will most likely improved the
performance of natural fibre composites by promoting better fibre to resin bonding.

ZUSAMMENFASSUNG: Naturfasern aus Hanf, Sisal, Jute und Kapok wurden mit zwei chemischen Metho-
den behandelt, um eine bessere Faser-Matrix-Anbindung in naturfaserverstärkten Verbundmaterialien zu
erreichen. Die Fasern wurden mit Natronlauge verschiedener Konzentration behandelt, um Oberflächenver-
unreinigungen zu entfernen und die Faserfeinstruktur zu modifizieren. Diese Fasern wurden dann mit und
ohne sauren Katalysator acetyliert, um die Hydrophilie der Fasern zu reduzieren und deren Witterungsbestän-
digkeit zu verbessern.
Die Auswirkung der chemischen Modifizierung wurde mit XRD, DSC, FT-IR und SEM untersucht. Nach
der Behandlung ist die Oberflächentopographie der Hanf-, Sisal- und Jutefasern sauber und rauh. Die Ober-
fläche der Kapokfasern blieb offensichtlich unverändert. Röntgenbeugungsmessungen zeigen einen leicht
erhöhten Kristallinitätsindex der mit niedrigen NaOH-Konzentrationen behandelten Fasern. Hohe NaOH-
Konzentrationen erniedrigen jedoch den Kristallinitätsindex. Die thermische Analyse der Fasern weist eben-
falls auf eine Erniedrigung des Kristallinitätsindexes mit zunehmender NaOH-Konzentrationen hin und daß
die Acetylierung bei höheren Temperaturen leichter abläuft. Mit der Alkalisierung und Acetylierung konnte
die Struktur der Naturfasern erfolgreich modifiziert werden, was deren Leistungsvermögen durch verbesserte
Faser-Harz-Bindung erhöhen dürfte.

1 Introduction generally referred to as ‘soft’ fibres to distinguish them


from the hard leaf fibres. Both leaf and bast fibres are
1.1 Plant fibres as reinforcement for composites multi-cellular with very small individual cells bonded
The increase in the application of plant fibres as reinfor- together1–3). Kapok is a single cell seed fibre with a wide
cement for polymeric substrates has been stimulated by air lumen.
the environmental cost of manufacturing energy-inten- The mechanical properties of plant fibres are largely
sive, synthetic fibres such as glass, carbon and Kevlarm. related to the amount of cellulose, which is closely asso-
However, whereas synthetic fibres can be produced with ciated with the crystallinity index of the fibre and the
engineered properties to suit particular applications this is micro-fibril angle with respect to the main fibre axis4).
not the case with naturally occurring plant fibres. Proper- Fibres with high crystallinity index and/or cellulose con-
ties of the cellulose fibres depend mainly on the nature of tent have been found to possess superior mechanical
the plant, locality in which it is grown, age of the plant properties. Sisal fibres with cellulose content of 67% and
and extraction method used. For instance, sisal is a hard micro-fibril angle of 10 – 22 8 have a tensile strength and
leaf fibre but jute and hemp are both bast fibres and are modulus of elasticity of 530 MPa and 9 – 22 GPa respec-

* Correspondence author.

Die Angewandte Makromolekulare Chemie 272 i WILEY-VCH Verlag GmbH, D-69451 Weinheim 1999 0003-3146/99/0112–0108$17.50+.50/0
Effect of chemical treatment on the properties of fibres 109

tively. On the other hand, coir fibre with a cellulose con- lulose II by a process known as mercerisation7, 8), which
tent of 43% and micro-fibril angle of 30 – 49 8 is reported in this paper is referred to as alkalisation. The reaction of
to have a tensile strength and modulus of elasticity of 106 sodium hydroxide with cellulose is thought to be as out-
MPa and 3 GPa respectively1, 3, 5). This variation in the lined in Eq. (1).
mechanical properties with increased microfibril angle
plays an important role in determining the mechanical Cell–OH + NaOH e Cell–O–Na+ + H2O
properties of fibre reinforced composites. In addition it is + [surface impurities] (1)
necessary to optimise fibre alignment parallel with the
direction of applied force to maximise tensile properties. It is worth pointing out that alkalisation de-polymerises
the native cellulose I molecular structure producing short
length crystallites. However, there seems to be varying
1.2 Fibre to matrix interfaces in natural fibre- interpretations of the term ‘mercerisation’. The standard
reinforced composites definition for mercerisation proposed by ASTM D1695 is
“the process of subjecting a vegetable fibre to the action
The performance and stability of fibre-reinforced compo- of a fairly concentrated aqueous solution of a strong base
site materials depends on the development of coherent so as to produce great swelling with resultant changes in
interfacial bonding between fibre and matrix. In natural the fine structure, dimension, morphology and mechani-
fibre-reinforced composites there is a lack of good interfa- cal properties” 9). ASTM D123-83a defines mercerised
cial adhesion between the hydrophilic cellulose fibres and yarn as “a cotton yarn which has been treated with a solu-
the hydrophobic resins due to their inherent incompatibil- tion of sodium hydroxide under conditions of concentra-
ity. Short, cellulose-based fibres will also tend to agglom- tion and temperature which effect a permanent or irrever-
erate making their use in reinforced composites less attrac- sible swelling of the cellulose” 10). In both definitions
tive. The presence of waxy substances on fibre surface neither the alkali concentration nor the treatment tem-
contributes immensely to ineffective fibre to resin bonding perature are mentioned.
and poor surface wetting is observed. Also the presence of Zeronian11) proposes another definition of mercerisa-
free water and hydroxyl groups, especially in the amor- tion, which is suitable for basic research and is more spe-
phous regions, worsens the ability of plant fibres to cific. Mercerised cellulose is “a sample of cellulose
develop adhesive characteristics with most binder materi- which has been treated with a solution of an alkali metal
als. High water and moisture absorption of the cellulose hydroxide of sufficient strength to cause essentially com-
fibres causes swelling and plasticising effect resulting in plete conversion of the crystal structure from cellulose I
dimensional instability and poor mechanical properties. to II”. It is reported that residual traces of cellulose I are
Plant fibres are also prone to micro-biological attack lead- found even when the strength of the alkali used in the
ing to weak fibres and reduction in their life span6). alkalisation treatment is considered optimum for conver-
Fibres with high cellulose content have also been found sion11). Atkins12) reports that alkalisation without tension
to contain a high crystallite content. These are the aggre- allows total conversion of cellulose I to cellulose II to
gates of cellulose blocks held together closely by the take place. However the application of tension to main-
strong intra-molecular hydrogen bonds which large mole- tain the even distribution of crystallites only allows par-
cules, for example dyes, are not able to penetrate unless tial conversion.
the cell wall is swollen. Fibres are, therefore, usually sub- Owolabi et al.13) used 50% NaOH on coconut fibres
jected to treatment such as alkalisation and acetylation, while Sreekala et al.4) and Geethamma et al.14) used 5%
with or without heat, to first bulk or swell the cell wall to NaOH to remove surface impurities on oil palm fibres
enable large chemical molecules to penetrate the crystal- and short coir fibres respectively. Bisanda and Ansell6)
line regions. applied a concentration of 0.5 N NaOH on sisal fibre,
which resulted in improvement in the mechanical proper-
ties of the sisal fibre. However, Murkhejee et al.15) found
1.3 Chemical treatment of natural fibres – that the use of more than 1% NaOH on cellulose fibres
alkalisation weakens the fibres resulting in poorer mechanical proper-
Natural fibres are chemically treated in order to remove ties. This last finding appears to be in perfect agreement
lignin-containing materials such as pectin, waxy sub- with the commonly held principle of alkalisation, which
stances and natural oils covering the external surface of states that caustic soda increases the amount of amor-
the fibre cell wall. This reveals the fibrils and gives a phous regions at the expense of crystallinity index.
rough surface topography to the fibre. Sodium hydroxide It is reported that alkalisation, both slack and with ten-
(NaOH) is the most commonly used chemical for bleach- sion increases the strength uniformity along the fibre
ing and/or cleaning the surface of plant fibres. It also length. This is attributed to an increase in strength at the
changes the fine structure of the native cellulose I to cel- weakest point in the fibre5, 8). Alkalisation also improves
110 L. Y. Mwaikambo, M. P. Ansell

accessibility of reactive sites to dyes and binding chemi- 1.5 Objectives


cals bringing about crystalline modification which To properly assess changes at the fibre surface and fine
involves fibril swelling and sometimes improves the crys- structure due to chemical treatment it is necessary to
talline packing order which has the advantage of provid- employ appropriate analytical characterisation methods.
ing more access to penetrating chemicals. The presence A combination of two or more characterisation techni-
of reactive sites and fibril swelling are prerequisite fac- ques allows a much more thorough investigation of the
tors to resin cross-linking inside the fibre. Cellulose- effect of chemical treatment on cellulose based fibres. In
based fibres absorb moisture causing both reversible and this study, therefore, four characterisation methods were
irreversible swelling. In composite products this can employed. Wide angle X-ray analysis (WAXS), differen-
result in undesirable dimensional changes. To arrest this tial scanning calorimetry (DSC), Fourier transform infra-
problem, the reinforcing cellulose fibres are subjected to red spectroscopy (FT-IR) and scanning electron micro-
certain modifications. These processes involves either the scopy (SEM) have been used to analyse the effect of mer-
stabilisation of the cell wall matrix to restrain swelling or cerisation and acetylation on the crystallinity index and
reduction of the hygroscopicity of the cell wall and bulk- thermal characteristics of four types of natural fibres.
ing of the cell wall polymers to maintain the wet volume
so that moisture does not cause any additional swelling6).
2 Experimental procedure
1.4 Chemical treatment of natural fibres – acetylation
2.1 Materials
Acetylation, which is another chemical treatment covered
Sisal and jute fibres used in this work were supplied by the
in this study, has been extensively applied to wood cellu-
Department of Materials Science and Engineering at the Uni-
lose to stabilise the cell wall, improving dimensional sta- versity of Bath. Kapok fibre was obtained from Morogoro in
bility and environmental degradation16) (Rowell, 1992). Tanzania whilst hemp was kindly supplied by the Hemcore
The process involves the soaking of the plant fibre in Company Ltd of United Kingdom. No specifications were
acetic anhydride with or without an acid catalyst as available regarding the physical characteristics of the sup-
shown in Eq. (2) and (3), respectively. plied fibres such as staple length, density, diameter and pro-
Acetylation with acid catalyst: cessing conditions. Sodium hydroxide pellets of 98%
strength, sulfuric acid with 99% strength and glacial acetic
acid were supplied as general laboratory reagents. Merck Ltd
of England supplied acetic anhydride with density of 1.08 g
cm–3 and boiling temperature of 140 8C. Chemicals were
diluted to the concentrations stipulated below.
Acetylation without acid catalyst:

2.2 Fibre preparation


Fibres were stored in a conditioning chamber containing a
saturated sodium nitrite solution whereby 85 g of the solute
Since acetic acid does not react sufficiently with cellu- was added to 100 cm3 of water. Assuming the room tempera-
lose, acetic anhydride is preferred instead17, 18). However, ture is 20 l 2 8C, the conditioning chamber with the solution
because acetic anhydride is not a good swelling agent for in it generates a relative humidity inside the chamber of
cellulose, in order to accelerate the reaction, cellulose approximately 65 l 2% relative humidity20). Fibres for the
materials are first soaked in acetic acid and subsequently acetylation treatment were soaked in cold distilled water for
48 h at 20 l 2 8C to remove any surface impurities that would
treated with acetic anhydride at higher temperatures for a
prevent effective action of the acetyl groups on the cellulose
period of between 1 and 3 h. The rate of reaction is much structure before they were conditioned. This process was
faster with fibres that have not been alkalised than with deemed unnecessary for fibres designated for alkali treat-
alkalised cellulose fibres. In natural fibre reinforced com- ment for reasons which will be explained in the alkali treat-
posite acetylation of the hydroxy group will swell the ment section. Fibre treatments are summarised in Tab. 1.
plant fibre cell wall, greatly reducing the hygroscopic nat-
ure of the cellulose fibre. This will consequently result in
dimensional stability of the composites, as any absorbed 2.3 Alkali treatment
water will not cause further swelling or shrinkage of the Kapok, sisal, jute and hemp were soaked in beakers contain-
composite material. While the acetylation treatment has ing caustic soda concentrations as shown in Tab. 1 and
long been used on textile goods, it is only recently that placed in a water bath controlled at 20 l 2 8C for 48 h. The
research work is being carried out to assess its usefulness fibres were then removed, washed with distilled water con-
in natural fibres for applications in composites4, 19). taining 1% acetic acid to neutralise excess sodium hydroxide
Effect of chemical treatment on the properties of fibres 111

Tab. 1. Specifications of chemical treatment of the fibres. intensities were recorded between 5 8 and 60 8 (2 h angle
range). The crystallinity index (Ic) was determined by using
Treatment Reaction Reaction temperature ( 8C) Eq. (4), where I(002) is the counter reading at peak intensity at
time a 2 h angle close to 22 8 and I(am) is the amorphous counter
Kapok Sisal Jute Hemp reading at a 2 h angle of about 18 8.
Untreated – – – – – Ic = (I(002) – I(am)) N 100/I(002) (4)
10% acetic acid (1 h)
+ acetic anhydride
in sulfuric acid for 5 min 20 20 20 20
1h 70 20 70 70 2.7 Infrared spectroscopy
3h 70 70 70 70 Infrared spectra were obtained using a Perkin Elmer FT-IR
10% acetic anhydride 1h 20 20 20 20 spectrometer model PARAGON 1000. About 2 mg of fibre
1h 70 70 70 70 was crushed into small particles in liquid nitrogen. The fibre
3h 70 70 70 70
particles were then mixed with KBr and pressed into a small
0.8% NaOH 48 h 20 20 20 20
2% NaOH 48 h 20 20 20 20 disc about 1 mm thick.
4% NaOH 48 h 20 20 20 20
6% NaOH 48 h 20 20 20 20
8% NaOH 48 h 20 20 20 20 2.8 Scanning electron microscopy (SEM)
30% NaOH 48 h 20 20 20 20
400% (4 M) NaOH 48 h 20 20 20 20 SEM micrographs of fibre surfaces and cross sections of
untreated and treated fibres were taken using a scanning
electron microscope Model JEOL 6310. Prior to SEM eva-
and then thoroughly rinsed with distilled water. The fibres luation, the samples were coated with gold by means of a
were then dried to remove free water and placed in a glass plasma sputtering apparatus.
container in a conditioning chamber.

3 Results
2.4 Acetylation
One set of the four fibres namely kapok, sisal, jute and hemp 3.1 Differential scanning calorimetry of mercerised
were treated in glacial acetic acid (Tab. 1) for 1 h at 20 l fibres
2 8C. It was further treated with acetic anhydride containing DSC analysis enables to identify the chemical activity
concentrated H2SO4 as a catalyst for 5 min. Fibres were then occurring in the fibre as as heat is applied and in this case
washed with distilled water and dried. A second set of fibres
it was possible to observe one endothermic peak at tem-
was treated in acetic anhydride without acid catalyst for 1 h
at 20 l 2 8C. The fibres were removed and washed with dis-
peratures between 70 – 100 and two exothermic peaks at
tilled water and dried. The third and fourth sets of fibres higher temperatures shown in Tab. 2. The first exothermic
were treated in acetic anhydride without acid catalyst in a peak reflects the stability of the fibres as a function of
water bath controlled at 70 8C for 1 h and 3 h, respectively. caustic soda concentration.
The fibres were then washed with distilled water and dried. Fig. 1 shows a sharp decreasing trend of the first
exothermic peak of temperature as the concentration of

2.5 Differential scanning calorimetry


Untreated and treated fibre samples weighing between 5 and
10 mg were placed in an aluminium capsule, sealed and
punctured to allow gases to escape during heating. A Du-
Pont DSC 2910 equipped with TA instruments was operated
in a dynamic mode with a heating scheme of 30 to 500 8C
and heating rate of 10 K min–1 in a nitrogen environment
purged at 25 mL min–1. The thermograms were analysed for
any changes in the thermal behaviour of the fibres.

2.6 X-ray diffraction (WAXS)


Untreated and treated fibres were mixed with a very small
amount of an adhesive material (Tragacanth BP), soaked in a
drop of distilled water and compressed into thin sheets and
dried. A wide-angle diffractometer equipped with a scintilla- Fig. 1. The effect of caustic soda on hemp, sisal, jute and
tion counter and a linear amplifier was used. The diffraction kapok as observed by the DSC method.
112 L. Y. Mwaikambo, M. P. Ansell

Tab. 2. Crystallinity index and first exothermic peak temperature as a function of concentration of alkali treatment (second exother-
mic peak temperature in brackets).

NaOH Hemp Sisal Jute Kapok


(%) Crystallinity Exotherm Crystallinity Exotherm Crystallinity Exotherm Crystallinity Exotherm
index (%) peaks ( 8C) index (%) peaks ( 8C) index (%) peaks ( 8C) index (%) peaks ( 8C)

0 87.87 357.00 70.9 365.31 78.47 369.48 45.75 359.50


(390.35) (444.63) (417.48) (402.87)
0.8 88.79 377.03 71.04 356.96 76.61 359.05 46.31 370.53
(389.31) (398.70) (417.48) (434.17)
2 88.75 369.50 68.93 351.75 83.10 354.88 49.5 352.79
(386.18) (412.26) (395.57) (402.00)
4 87.61 364.27 74.66 353.24 83.06 346.55 48.73 367.00
(389.31) (387.04) (375.77) (406.00)
6 86.54 363.22 79.30 347.57 82.37 349.66 48.02 341.31
(382.00) (394.52) – (416.43)
8 87.70 362.18 75.11 350.74 82.50 351.75 58.42 353.96
(397.65) (393.30) (444.60) (404.96)
30 89.77 359.05 78.77 350.74 82.50 342.36 62.41 343.40
(389.31) (396.61) (391.39) (409.13)
4M 81.34 363.22 73.51 348.23 78.32 349.66 53.74 359.05
(414.35) (374.52) – (385.13)

the caustic soda is increased from 0.8% to 6%. The peak


temperature then slowly decreases up to 30% NaOH
beyond which a slight increase is observed. The informa-
tion is also summarised in Tab. 2.
Comparing the untreated fibres, hemp is the least resis-
tant to thermal treatment. However, once treated with
caustic soda hemp fibre gives higher thermal resistance
than the rest of the alkali treated fibres except at 4%
NaOH concentration where kapok fibre gives higher ther-
mal resistance. Sisal and jute show a drastic decrease in
thermal resistance when subjected to concentrations of
0.8% and 4% NaOH and then the resistance rises slightly
between 4 and 8% NaOH concentration after which it
drops. With the exception of sisal, the thermal resistance
of other fibres rises above 30% NaOH concentration.
Fig. 2. The effect of acetylation on hemp (ANHDH701), sisal
(ANHDS701), jute (ANHDJ701) and kapok (ANHDK701) as
observed by the DSC method.
3.2 Differential scanning calorimetry of acetylated
fibres
The thermal characteristics of acetylated hemp, sisal, jute
and kapok fibres are shown in Fig. 2. 3.3 X-ray diffraction of mercerised fibres
All the fibres show an endothermic peak at around As the concentration of caustic soda increases, X-ray dif-
80 8C, which is due to water, desorption. The breaking fraction results for alkalised fibres (Fig. 3 and Tab. 2)
down of the acetyl group causes the second endothermic show an overall increase in crystallinity index (Eq. (4))
peak observed in jute fibre. Beyond this peak nearly all for sisal, jute and kapok fibre while that of hemp fibre
the fibres show an exothermic peak between 380 8C and decreases. Hemp fibre has the highest crystallinity index
400 8C while kapok fibre has two minor exothermic peaks at any of the caustic soda concentrations followed by
between 300 8C and 350 8C and one major peak at around jute, sisal and kapok respectively. The effect of alkalisa-
400 8C. On application of heat kapok fibre is seen to be tion on the fibres was not the same at each of the applied
less thermally stable followed by sisal then hemp. The caustic soda concentrations. For example, at 0.8% NaOH,
exothermic peak seen in jute, hemp and sisal is more reg- jute fibre has a lower crystallinity index than the
ular in acetylated fibres. untreated jute fibre whereas a slight increase in crystalli-
Effect of chemical treatment on the properties of fibres 113

Tab. 3. Infrared transmittance peaks (cm–1) of acetylated fibres


relative to untreated fibres.

Bond type Hemp Sisal Jute Kapok

C1H stretchinga) 3448 3423 3471 3406


C1H stretching 2916 2920 2920 2920
Carboxylic anhydride 1638 1740 1743 1740
C1H bending 1383 1384 1384 1374
C1H bending – 1250 1248 1244
C1C stretching 1059 1509 1059 1057

a)
With intermolecular hydrogen bonding.

1743 cm–1 in sisal, jute and hemp and kapok fibres is due
to the bonded acetyl group. The peak at 1638 cm–1 in
hemp fibre is due to removal of unsaturated C2C stretch-
ing present in traces of oils.
Fig. 3. Crystallinity index versus caustic soda concentration Similar observations have been reported in earlier work
for untreated and treated hemp, sisal, jute and kapok fibres on acetylation of the wood cell wall21–23). The increase in
measured by WAXS.
absorbency in the region between 1000 – 1500 cm–1 bands
shows the increase in O1H stretching, indicating that
nity index is observed in hemp, sisal and kapok at the there has been a reduction in the number of hydroxy
same alkali strength in contrast with the untreated fibres. groups at the 3400 to 3500 cm–1 band.
At 2% NaOH sisal, hemp and kapok show increases in
crystallinity index while jute fibre shows a decrease in
the crystallinity index. Hemp and kapok fibres have their 3.5 Scanning electron microscopy of alkalised fibres
highest crystallinity index at 30% NaOH, sisal at 6% Following alkalisation the surface topography of jute,
NaOH, and jute at 4% NaOH. sisal and hemp fibres is rougher than before treatment
(Fig. 4 – 6).
Sisal, hemp and jute comprise bundles of individual
3.4 FT-IR analysis of acetylated fibres cells that have been bound together by lignin-rich, weak
The grafting of acetyl groups to fibre cell walls is of con- inter-molecular bonds. Sisal fibres are discontinuous,
siderable importance to this work. FT-IR reveals the comprising short lengths joined together to end, whereas
extent of grafting. The characteristic peaks observed are hemp and jute fibres are continuous. Fig. 4 (b) show
summarised in Tab. 3. The peak observed at 3440 cm–1 in ridges on the surface of clean jute fibre after alkali treat-
untreated fibres indicates the presence of intermolecular ment. Hemp fibre shows partly separated individual cells
hydrogen bonding and tends to shift to higher absorbency before alkalisation, Fig. 6 (a). Alkali treated hemp fibre
values in acetylated fibres, e. g. 3480 cm–1 in sisal, jute looks cleaner and fibre bundles are more separated, with
and hemp fibres but remains unchanged at around 3400 a highly serrated surface. The surface of kapok fibres
cm–1 in kapok fibre. The increase in peak intensity at appears to be unaffected by alkalisation (Fig. 7).

Fig. 4. (a) Untreated jute fibre and (b) alkali treated jute fibre.
114 L. Y. Mwaikambo, M. P. Ansell

Fig. 5. (a) Untreated sisal fibre (b) and alkali treated sisal fibre.

Fig. 6. (a) Untreated hemp fibre and (b) alkali treated hemp fibre.

Fig. 7. (a) Untreated kapok fibre and (b) alkali treated kapok fibre.

4 Discussion ful tool to determine the drop in crystallinity index and


decomposition of plant fibre cellulose. The results
4.1 Differential scanning calorimetry obtained in this work using the first exothermic DSC
Caustic soda treatment of plant fibres has, therefore, dual peak (decomposition of the cellulose) corresponds well
advantages; it removes the fibre surface impurities with the results obtained using the second and stronger
(Fig. 4 – 7) and more importantly, modifies the crystallites endothermic peak (reduction in the crystallites) to assess
of the cellulose (Fig. 3). The process also swells the fibre the thermal degradation of crystallites in plant fibres.
to enhance the crystallite order and increase chemical Shenouda8) and Nguyen et al.24) have made similar obser-
uptake. The DSC technique is reported24) to be a very use- vations with respect to the second peak.
Effect of chemical treatment on the properties of fibres 115

The sharp decrease in the decomposition temperature Sreekala4) and Hill et al.29) studied the acetylation of coir
of the alkalised plant fibres between 0.8% and 8% NaOH fibre and found similar results. Acetylation of plant fibres
concentration is an indication of the increase in amor- reduces the hygroscopic nature of the cell wall and the
phous cellulose, known to have poor thermal resistance, incorporation of acetylated fibres into plastics enhances
and a decrease in cellulose crystallite length. This view is weather resistance, thermal resistance and dimensional
shared by several researchers6, 25). It is observed in Fig. 1 stability of the composites16, 18, 30).
that more than 8% caustic soda renders the converted cel-
lulose structure slightly stable to thermal degradation
(Fig. 1). Alkali treated hemp fibre is more crystalline 5 Conclusions
than sisal, jute and kapok (Tab. 2 and Fig. 3) and was – Alkalisation of plant fibres effectively changes the sur-
found to be more stable to thermal degradation as mea- face topography of the fibres and their crystallo-
sured by the DSC. graphic structure. However care must be exercised in
Some researchers report that alkalised plant fibres used selecting the concentration of caustic soda for alkalisa-
as reinforcement in the manufacture of composites tion as results show that some fibres at certain NaOH
improve mechanical properties in comparison with non- concentration have reduced thermal resistance as elu-
alkalised fibres6, 26). Mwaikambo and Bisanda27) found cidated by the DSC method.
that the application of 5% sodium hydroxide to cotton/ – It is believed that the increase in the crystallinity index
kapok fabric for reinforcement of unsaturated polyester obtained by X-ray diffraction is in actual fact an
resin increased tensile strength but decreased modulus of increase of the order of the crystallite packing rather
elasticty and impact strength. Aboul-Fadl et al.25) found than in increase in the intrinsic crystallinity index.
that there was a decrease in the breaking strength and – It is essential therefore to use several complementary
tenacity of alkalised Pima S-5, Giza 76, and Giza 77 cot- techniques when studying the fine structure of natural
ton species while alkalised Giza 75, Giza 80 Dendara, fibres to confirm trends and that the application of the
Deltapine Smooth Leaf produced higher breaking and DSC technique probably gives a better analysis of the
tenacity values. This implies that the alkalisation of plant fine structure of the plant fibres than the X-ray method
fibres can have different effects on the mechanical prop- alone.
erties of fibres and also composite materials reinforced – The removal of surface impurities on plant fibres may
with these fibres. be an advantage for fibre to matrix adhesion as it may
facilitate both mechanical interlocking and the bond-
ing reaction due to the exposure of the hydroxyl
4.2 X-ray diffraction groups to chemicals such as resins and dyes.
X-ray results for alkalisation which show an overall
I am grateful to the Sokoine University of Agriculture for
increase in the ‘crystallinity’ index indicate improvement availing financial support in a form of scholarship under the
in the order of the crystallites as the cell wall thickens NORAD TAN 91 programme financed by the Norwegian Gov-
upon alkali treatment. Alkali treatment is reported to ernment. I wish also to thank the Department of Materials
reduce the proportion of crystalline material present in Science and Engineering at the University of Bath for providing
travel funds to attend the symposium.
plant fibres, as observed by several researchers5, 25, 28). It is
therefore difficult to reconcile the results of this work.
However, since alkalisation with and without tension 1) R. D. Preston, “Observed fine structure in plant fibres”, in:
increases the crystallite packing order, it is therefore logi- Fibre structure, J. W. S. Hearle, R. H. Peters (Eds.), Butter-
cal to deduce that the order of the crystallites improves worth, London (1963), Chapter 7
2)
rather than the crystallinity index increasing. J. W. S. Hearle, “The development of ideas of fine structure”,
in: Fibre structure, J. W. S Hearle, R. H. Peters (Eds.), Butter-
worth, London (1963), Chapter 6
3) L. Hegbom, B. Ultne, Microscopy studies of some non-wood
4.3 FT-IR analysis raw materials for the pulp and paper industry, Chemical and
The FT-IR results in Tab. 3 indicate that some chemical Process Engineering for Development – A challenge for the
21st century (1990)
reactions occurring during acetylation of the fibres and 4) M. S. Sreekala, M. G. Kumaran, S. Thomas, J. Appl. Polym.
the presence of a peak in all the fibres at 1740 cm–1 is Sci. 66 (1997) 821
caused by the reaction of the ester groups present at 1734 5) J. W. S. Hearle, “Structure properties and uses”, in: Fibre
cm–1 in the untreated fibres with the acetyl groups structure, J. W. S Hearle, R. H. Peters (Eds.), Butterworth,
observed at the former peak. Similarly, the reduction of London (1963), p. 621
6) E. T. N. Bisanda, M. P. Ansell, J. Mater. Sci. 27 (1992) 1690
the intermolecular hydrogen bonding between 3406 and 7) D. J. Johnson, “High-temperature stable and high-perfor-
3471 cm–1 confirms the grafting of the acetyl groups on mance fibres”, in: Applied Fibre Science, Vol. 3, F. Happey
the cellulose structure thus replacing the hydroxyl groups. (Ed.), Academic Press, London, New York (1979), Chapter 3
116 L. Y. Mwaikambo, M. P. Ansell

8) S. G. Shenouda, “The structure of cotton cellulose”, in: 19) H. P. S Abdul Khalil, “Acetylated plant fibre reinforced com-
Applied Fibre Science, Vol. 3, F. Happey (Ed.), Academic posites”, PhD Thesis, School of Agricultural and Forest
Press, London, New York (1979), p. 275 Sciences, University of Wales, Bangor, Gwynedd, United
9) ASTM D 1695, Annual Book of ASTM Standards, section Kingdom (1999)
15, Vol. 15.04; Soap, Polishes, Cellulose, Leather, Resilient 20) WIRA, Textile Data Book, Tab. E2 (1973)
Floor Coverings (1983) 21) C. Clemsons, R. A. Young, R. M. Rowell, Wood Fibre Sci.
10) ASTM D 123-83a, Annual Book of ASTM Standards, section 24 (1992) 353
7, Vol. 07.01; Textiles – Yarns, Fabrics, General Tests Meth- 22) V. C. Mallari, K. Fukuda, N. Morohoshi, T. Haraguchi,
ods (1983) Mokuzai Gakkaishi 32(2) (1990 or 1986??) 139
11) S. H. Zeronian, “Intra-crystalline swelling of cellulose”, in 23) R. M. Rowell, R. Simonson, S. Hess, D. V. Placket, D . Cron-
“Cellulose chemistry and its applications”, T. P. Nevell, S. H. shaw, E. Dunningham, Wood Fibre Sci. 26 (1994) 11
Zeronian (Eds.), E. Horwood/Halsted Press, Chichester/New 24) T. Nguyen, E. Zavarin, E. M. Barrall II, J. Macromol. Sci.,
York (1985), p. 159 Rev. Macromol. Chem. Phys. C20 (1981) 1
12) E. Atkins, “Polysaccharides: Biomolecular shape and struc- 25) S. M. Aboul-Fadl, S. H. Zeronian, M. M. Kamal, M. S. Kim,
ture”, in: Applied Fibre Science, Vol. 3, F Happey (Ed.), Aca- M. S. Ellison, Text. Res. J. 55 (1985) 461
demic Press, London, New York (1979), Chapter 8 26) N. E. Marcovich, M. M. Reboredo, M. I. Aranguren, J. Appl.
13) O. Owolabi, T. Czvikovszky, I. Kovacs, J. Appl. Polym. Sci. Poym. Sci. 70 (1998) 2121
30 (1985) 1827 27) L. Y. Mwaikambo, E. T. N. Bisanda, The performance of cot-
14) V. G. Geethamma, R. Joseph, S. Thomas, J. Appl. Polym. ton – kapok fabric – polyester composites, Vol. 18 (3)
Sci. 55 (1995) 583 (1999) 181
15) A. Murkhejee, P. K. Ganguly, D. Sur, J. Text. Inst. 84 (1993) 28) W. E. Morton, J. W. S. Hearle, “An introduction to fibre
348 structure”, in: Physical Properties of Textile Fibres, Heine-
16) R. M. Rowell, “Property enhancement of wood composites”, mann (for the Textile Institute), London (1975), p. 1
in: Composite applications – the role of matrix, fibre and 29) C. A. S. Hill, H. P. S. Abdul Khalil, M. D. Hale, “A study of
interface”, R. M. Rowell, T. Vigo, B. Kinzig (Eds.), VCH the potential of acetylation to improve the properties of plant
Publishers, New York (1992), Chapter 14 fibre”, in: Industrial Crops and Products (1998), Vol. 8, p. 53
17) Please contact the authors for further details 30) E. Obataya, J. Wood Sci. 45(2) (1999) 106
18) R. W. Moncrieff, “Cellulose acetate”, in: Man-made fibres,
Newnes-Butterworth, London (1975), p. 232

You might also like