Effect of Chemical Treatments On Properties of Raffia Palm Fibers - Fadele
Effect of Chemical Treatments On Properties of Raffia Palm Fibers - Fadele
https://doi.org/10.1007/s10570-019-02764-8 (0123456789().,-volV)
( 01234567
89().,-volV)
ORIGINAL RESEARCH
Abstract Chemical treatment is used to improve calorimetry, universal testing machine and Fourier
interfacial bonding of natural fibers to polymers to transform infrared spectroscopy. Chemical treatment
produce natural fiber reinforced polymer matrix reduced the tensile strength of the fibers. Some non-
composites with enhanced mechanical properties. In cellulosic components were partially or fully removed
the present study, raffia palm (Raphia farinifera) fibers as a result of chemical treatment. There was an
intended for use in composite manufacture was increase in the degradation temperature of the fibers
subjected to chemical treatments with NaOH and due to chemical treatment. Water absorption capacity
H2SO4 solutions. Morphological, thermal, tensile, of the fibers decreased as a results of chemical
physical and structural changes in the fibers before treatment, which will be beneficial in composite
and after treatment were investigated using scanning synthesis.
electron microscopy, differential scanning
M. Soleimani L. G. Tabil
Department of Chemical and Biological Engineering,
College of Engineering, University of Saskatchewan, 57
Campus Drive, Saskatoon, SK S7N 5A9, Canada
123
Cellulose
Graphic abstract
Keywords Raffia palm fiber Mechanical thermoplastic and thermoset matrices have been
properties Thermal properties Fourier transform embraced by European car makers and suppliers for
infrared spectroscopy door panels, seat backs, headliners, package trays,
dashboards, and other interior parts (Holbery and
Houston 2006).
In spite of the aforementioned attractive proper-
Introduction ties of NFs, their use in several structural applica-
tions are limited. They have high variability in
The mechanical properties of polymers are inadequate properties (Chinga-carrasco et al. 2011); these
for many structural purposes; their strength and properties depend strongly on age (maturity level),
stiffness are lower than those of ceramics and metals climate, weather and location. They are hydrophilic,
(Santhosh et al. 2014). As such, they are often which results in poor interfacial bonding between
reinforced with strong but expensive synthetic fibers, the polymer matrix and fiber, and swelling of NF-
whiskers or particles to achieve improved mechanical reinforced PMCs thereby leading to poor mechan-
properties and widen their structural applications ical performance. To overcome the latter problems,
(Thiruchitrambalam et al. 2010). However, polymer NFs are often chemically treated to modify the
composites made from synthetic reinforcements pose surface properties (Bledzki and Gassan 1999;
severe environmental pollution problems as they are Benyahia and Merrouche 2014). Chemical surface
not readily recyclable (Elenga et al. 2009). Hence, the treatment of NFs cleans the fibers by removing
increasing interest in the use of plant-based natural pectin, waxes, proteins, natural oils and other
fibers (NFs) for making polymer matrix composites impurities; it also increases their surface roughness
(PMCs). NFs are lightweight, inexpensive, renewable which is vital in developing good bonding to
and biodegradable. Some NFs have mechanical prop- polymer matrices. Good bonding in composites
erties comparable to those of some synthetic fibers enhance load transfer from the fibers to the matrix
(Mohammed et al. 2015). NF composites with via interfacial shear stresses.
123
Cellulose
Raffia palm fibers (RPFs) traditionally find appli- NaOH concentration and treatment time and temper-
cations in products such as ropes, yarns, dresses, ature was based on the results of several tests
carpets, mats, ligatures as well as in reinforcements for conducted in-house to optimize the cellulose content
bricks, cements and concretes (Sandy and Bacon of RPF through alkalization treatment. Similarly,
2001; Elenga et al. 2009; Odera et al. 2015). Studies by another set of ground RPFs was immersed in 0.6
Elenga et al. (2009) and Sandy and Bacon (2001) molar aqueous solution of H2SO4 (with a solution to
showed that RPFs have good potential as reinforce- fiber ratio of 10 ml to 1 g) at 100 °C for 2 h. Again,
ment for polymers in composites manufacture. Nev- this treatment gave the largest amount of cellulose
ertheless, before they are used for this application, it is without causing lignin hydrolysis from a set of tests
imperative to understand how chemical surface mod- conducted in-house. After the immersion, the fibers
ification affects their properties. This will help in were washed in laboratory water and then in distilled
optimizing the properties of polymer composites in water. Drying of the washed samples was carried out
which they will be used. Currently, there is dearth of in an air furnace at 60 °C for 24 h. The particle density
information regarding how various chemical treat- of the washed and dried fibers was measured at room
ments adopted for other NFs affect the properties of temperature using a gas pycnometer (Quanta-chrome
fiber reinforced polymers. The purpose of this study is Instruments, Boynton Beach, FL). The average parti-
to investigate the effects of sodium hydroxide and cle densities obtained for untreated, NaOH treated,
sulphuric acid treatments on the properties of RPFs. and H2SO4 treated fiber were 1.50 ± 0.01 g/cm3,
Changes in properties can be used for forensic 1.53 ± 0.01 g/cm3 and 1.52 ± 0.01 g/cm3, respec-
comparison of RPFs to determine fiber origin and tively. A slight positive increase was observed in the
counterfeit as well as ranking the suitability of a given densities of the treated fibers in comparison to
batch of RPFs for making biocomposites. untreated fiber. This increase is primarily due to the
densification of the treated fibers cell wall (Aziz and
Ansell 2004; Hashim et al. 2017).
Materials and methods
Chemical composition
Material
Chemical composition of untreated and chemically
The RPFs used in the present study were obtained treated RPFs were determined using a Ankom 200
from the rainforest region of Nigeria in partially Fiber AnalyzerTM (ANKOM Technology, Macedon,
processed form. Remnant binders sticking to the fibers NY) in accordance with the detergent fiber analysis
were manually removed with 3M ProGrade Precision method developed by Peter Van Soest (Soleimani
3.7 9 9 Sanding Sheets 220 grit papers (3MTM, USA). et al. 2008; Jin et al. 2013). Ankom 200 Method 5
After this, the fibers were soaked in 2% formulated (Ankom Technology 2000), Method 6 (Ankom Tech-
detergent solution for 10 min and cleaned to remove nology 2011) and Method 8 (Ankom Technology
oily substances and other impurities on the fiber 2013) procedures were used to obtain the standard acid
surface. After cleaning, the fibers were dried in an air detergent fiber (ADF), neutral detergent fiber (NDF)
oven at 70 °C for 24 h and subsequently cooled to and acid detergent lignin (ADL) data, respectively.
room temperature before further testing. Fibers used 0.5 g of ground untreated and treated dried fibers was
for chemical composition, density determination, used for each analysis. The difference between ADF
thermal analysis, X-ray diffraction and FTIR analyses and ADL readings gave the cellulose content, the
were ground with Retsch cutting mill SM100 (Retsch, hemicellulose content was based on the difference
Retsch-Allee, Haan) and sieved using a 1 mm screen. between the NDF and ADF readings, while ADL
reading provided the lignin content. The data reported
Chemical treatment in this study for each test are the average values of
three measurements (dry-matter basis).
Some ground RPFs were immersed in 10% w/v
aqueous NaOH solution (with a solution to fiber ratio
of 10 ml to 1 g) at 60 °C for 5 h. The selection of the
123
Cellulose
Colour measurements of RPFs mounted specimens were initially polish using 320
(46 lm), 500 (30 lm) and 1200 (15 lm) SiC grit
The natural colour of NF is due to the presence of non- emery papers and followed by fine polishing using
cellulosic constituents (Reddy and Yang 2005); this 2000 (10 lm) and 4000 (5 lm) SiC grit emery papers.
may be affected by chemical treatments thereby Finally, they were polished using 1 lm MD-Nap cloth
leading to improper identification by visual examina- with 1 lm MD-Nap suspension. The polished speci-
tion. Also, since colour is an important property in a mens were gold-coated using an Edwards S150B
product, the colour analysis of the fibers before and sputter coater (BOC Edwards, Burgess Hill, UK) and
after chemical treatments is necessary. Colour mea- imaged in the SEM using an accelerating voltage of
surements were carried out to ascertain the difference 3 kV.
in colour between untreated and treated RPFs. A
HunterLabTM ColorFlex EZ spectrophotometer (Hun- FTIR spectroscopy
ter Associates Laboratory Inc., Reston, VA) with a
port size of 30 mm in diameter was used for the Fourier transform infrared (FTIR) spectroscopy was
measurement. The test was conducted in accordance used to characterize the chemical structures of
with ASTM E1164-12 standard (ASTM E1164-12 untreated and treated RPF in order to determine any
2017) and the determination of whiteness and yellow- structural changes caused by alkaline and acid treat-
ness indices was done according to ASTM E313-15 ments. A Renishaw Raman inVia Reflex Microscope
standard (ASTM E313-15 2015). Approximately 2 g (Renishaw Inc, West Dundee, IL) coupled to a Smiths
of ground untreated and treated dried fibers was used IllumminantIR IITM spectrometer equipped with an all
to determine the Hunter L*, a* and b* coordinates. reflective objective (ARO) and a diamond attenuated
The value of L* coordinate represents the whiteness total reflection (ATR) objective was used for both
component (0 = black, and 100 = white); the a* analyses. The FTIR data were acquired at a resolution
coordinate represents greenness to redness ð a = of 4 cm-1 and 512 scans per sample.
green, and þ a = red); and b* coordinate represents
blueness to yellowness ð b = blue, and þ b = yel- Tensile test
low). Therefore, an increase in L*, a* and b* denotes
more white, red and yellow colours, respectively. The The effect of chemical treatment on the tensile
variation of colour (DE) was estimated in comparison properties of unground RPF was investigated using a
to the raw fiber using Eq. 1 (Soleimani et al. 2008). 5 kN capacity Model 3366 InstronÒ universal testing
h 2 2 2 i0:5 machine (Instron Corp., Norwood, MA). The test was
DE ¼ Lt L þ at a þ bt b ð1Þ conducted at room temperature and a relative humidity
of 35% according to ASTM D3822-14 standard
where the coordinates with subscript t are for treated (ASTM D3822M-14 2014). The crosshead speed used
RPFs, while those without subscripts are for untreated was 1 mm/min and the fiber length used was 45 mm.
RPFs. The whiteness (L*) and chromacity coordinates The diameter of each tested fiber was measured at five
(a* and b*) for each sample were measured in 10 locations along the length and the average diameter
replicates and the average values of the results are obtained was then used in area calculation for tensile
presented. strength determination. To avoid slip from or damage
by the sample grip system of the testing machine, the
Microstructure fibers were covered with tissue papers at both ends and
a cellophane tape was used to fasten them to the fibers.
The microstructures of untreated and chemically
treated RPFs were examined using a Hitachi FE-
SEM SU8010 (Hitachi High-Technologies Corp.,
Tokyo, Japan) scanning electron microscope (SEM).
For the cross-sectional imaging, RPF specimens were
cold mounted using acrylic resin to produce a smooth
Fig. 1 A schematic drawing of a single raffia palm fiber
fiber cross sectional surface for imaging. The cold- prepared for tensile test
123
Cellulose
Figure 1 shows a drawing of how each fiber was The diffractometer was operated at 40 kV and 44 mA.
prepared for testing. The measurements were carried out on the Multipur-
Ten fibers were tested for each group of sample (i.e. pose Attachment, with parafocusing mode. A Kb filter
untreated and chemically treated) and the reported (Ni foils) was placed at the receiving end. The
tensile data are the averages of the ten measurements. diffraction patterns were measured from 5° to 50°,
The tensile test data were further analyzed statistically with a scan rate of 0.5° per minute (step size: 0.02°),
by one-way analysis of variance (ANOVA) to deter- for all the samples.
mine if the tensile properties (tensile strength and Assignment of Miller indices to the diffraction
percent elongation) of the fibers were significantly peaks was done with the assistance of Mercury
affected by chemical treatment. The SigmaPlot soft- software (version 4.1.0) program (Bruno et al. 2002;
ware for Windows Version 13 (Systat Software Inc., Macrae et al. 2008; French 2014). The crystal
San Jose, CA) was used in the ANOVA test. information file (.cif) for cellulose Ib, which was
obtained from the Cambridge Crystallographic Data-
Water absorption base with Reference code JINROO01 (cellulose Ib),
was used to simulate representative spectra of cellu-
Water absorption test was carried to determine how lose Ib. The selection of cellulose 1b crystal file for the
chemical treatment affected the water absorption present analysis was because it is the dominant
capability of dry unground RPFs. Several fibers, each cellulose polymorph in land plants (Nishiyama et al.
weighing approximately 0.50 g, were first dried for 2002; Lee et al. 2015; Nam et al. 2016). The
60 h at 70 °C, cooled to room temperature and then parameters used for the simulated XRD patterns were
weighed before immersing in distilled water at room Cu–Ka wavelength of 0.154 nm (1.54056 Å) and full
temperature. The water absorption test was carried out width at half maximum (FWHM) values of 3.0°, 3.5°
based on similar studies (Ramadevi et al. 2012; and 4.0°. The three FWHM values were chosen to
Ronald Aseer et al. 2013; Nayak and Mohanty 2018) determine which one yields a XRD pattern that best
found in literatures on natural fibers by subsequently fits the experimental data for the untreated raffia palm
immersing the dried fibers in a water bath containing fiber. The simulated spectra obtained are shown in
distilled water maintained at room temperature. At Fig. 2. The two crystalline diffraction peaks (1 - 1 0)
different times, five fibers were removed from the and (1 1 0) at 2h * 15.1° and 2h * 16.6°, respec-
water bath, cleaned to remove any water on the fiber tively, coalesce into a single peak at 2h * 15.7° as
surface and reweighed within 30 s to minimize FWHM value increased from 3.0 to 4.0 (French and
evaporation effect. The weight data were collected Santiago Cintrón 2013).
as a function of immersion time until saturation was
reached. The water intake content (WC) of the fibers
(in wt%) was calculated using Eq. 2 (Nayak and
Mohanty 2018).
Wt W0
WCt ðwt%Þ ¼ 100 ð2Þ
W0
where Wt and W0 denote the weight of the fibers after a
specific time, t, and the dry weight of the fibers
respectively.
123
Cellulose
Thermal analysis which could not be quantified using the Ankom 200
Fiber AnalyzerTM. The cellulose content of untreated
The effect of chemical treatment on thermal behavior RPFs is higher than those of wheat straw (43.2%)
of RPFs was studied using both differential scanning (Alemdar and Sain 2008), wheat straw (38–40.8%)
calorimetry (DSC) and thermogravimetric analysis (Bjerre et al. 1996); coconut coir (32.7%) (Jústiz-
(TGA). DSC analysis was done using a model 2910 Smith et al. 2008), banana (43.5%) (Jústiz-Smith et al.
V4.4E TA Instruments modulated differential scan- 2008), and alfa grass (46 ± 3%) (Mabrouk et al. 2012)
ning calorimeter (MDSC) (TA Instruments, New which have been used to fabricate polymer compos-
Castle, DE). Both the untreated and chemically treated ites. Mechanical properties of NF depend strongly on
RPFs were oven-dried at 60 °C for 24 h before the their cellulose content since it provides strength and
test. Each scan was performed under argon gas in an stability to the fibers (Bledzki and Gassan 1999). A
open aluminum pan from room temperature to 400 °C close look at the data in Table 1 shows that neither
at a heating rate of 5 °C/min. Approximately 10 mg of alkaline treatment nor sulfuric acid treatment could
ground untreated and treated RPFs were used and the delignify RPFs completely. Relative to the untreated
measurement was repeated twice to ensure fiber, there was a 22% increase in the cellulose
reproducibility. content, a slight increase (* 3%) in hemicellulose
TGA test was conducted using a TA instruments content, while there was * 8% decrease in lignin
TGA Q5000 operated between room temperature and content after NaOH treatment. On the other hand, after
600 °C at a heating rate of 10 °C/min with a nitrogen H2SO4 treatment, there was a 22% increase in
gas flow rate of 25 mL/min. The main purpose was to cellulose content, roughly 32% increase in lignin
compare the desorption and degradation behaviour of content while the hemicellulose content reduced by
untreated raffia palm fibers with that of chemically 97%. Changes in cellulose, hemicellulose and lignin
treated fibers. About 20 mg of each ground fiber contents of NF subjected to chemical treatments have
sample was used. The mass loss rate (in wt%/°C) for been well documented in the literature (Arsène et al.
each sample was derived from derivative TG (DTG) 2007; Vardhini et al. 2016; Soleimani et al. 2017). The
data. amount of cellulose, hemicellulose and lignin
removed by a chemical treatment depends on the
nature of the fiber, concentration and temperature of
Results and discussion the chemical and the length of treatment time.
Vardhini et al. (Vardhini et al. 2016) reported a 20%
Chemical composition of raffia palm fibers increase in cellulose content and a 50% decrease in
hemicellulose content of banana fibers after treatment
The chemical composition (dry matter basis) of RPFs in 11 g/L solution of NaOH at 90 °C for 2.5 h.
before and after chemical treatments in wt% are Soleimani et al. (Soleimani et al. 2017) reported about
presented in Table 1. Untreated RPFs comprise cellu- 70% increase in cellulose content and about 95%
lose (53 wt%), hemicellulose (13 wt%) and lignin (24 decrease in hemicellulose content after pretreating oat
wt%). The remaining 10 wt% is attributed to other hull fibers with 0.1 N of H2SO4 at 130 °C for 40 min.
components (e.g. pectin, wax, protein, oil, and ash)
Table 1 Chemical composition of untreated and chemically treated raffia palm fibers (dry matter basis)
Fiber Soaking time Cellulose Hemicellulose Lignin Others Dry matter Moisture
(h) (wt%) (wt%) (wt%) (wt%) (wt%) (wt%)
123
Cellulose
However, Arsène et al. (Arsène et al. 2007) reported responsible for the natural colours of NF (Reddy and
that the hemicellulose content of banana trunk fibers Yang 2007; Rokbi et al. 2011).
mercerized by boiling in 5% Ca(OH)2 for 1 h followed
by rinsing in deionized water and drying remained Microstructure
practically the same as that of the untreated fibers.
Figure 3 shows typical SEM micrographs of the
Colour of raffia palm fiber transverse section of untreated and treated RPFs.
After chemical treatment, the microstructure of cross-
One of the physical changes accompanying the sections of alkaline (Fig. 3b) and acid (Fig. 3c) treated
chemical treatment of RPFs is a change in colour of RPFs are essentially the same as that of the untreated
the fibers. The results of colour measurements which fiber (Fig. 3a). The surfaces of fibers treated with
were obtained using the Hunter L*, a* and b* NaOH and H2SO4 as shown in Fig. 4, are much
coordinates are presented in Table 2. The value of cleaner, rougher and porous than those of untreated
L* coordinate represents the whiteness component fibers, which is an indication that these chemicals
(0 = black, and 100 = white); the a* coordinate removed wax, protein, pectin, oil and other impurities
represents greenness to redness ð a = green, and from the surfaces of the fibers. It was reported that the
þ a = red); and b coordinate represents blueness to use of chemical treatment led to the removal of non-
yellowness ð b = blue, and þ b = yellow). The cellulosic components and also resulted in changes in
results show that the fiber became 23% darker after both surface chemistry and thermal properties of NFs
alkaline treatment, which is consistent with the results (Gonzalez et al. 1999; Cicala et al. 2010). The removal
obtained by Elenga et al. (Elenga et al. 2013) who of these non-cellulosic materials as well as the
reported that Raphia textilis fiber became darker by increased surface roughness is expected to promote
13% after alkaline treatment. Also, the redness and strong bonding between the fibers and polymer
yellowness of the fibers were reduced respectively by matrices when used in composite manufacture (Kabir
27% and 14% in comparison with the untreated fiber. et al. 2012).
After acid treatment, the fibers became 28% darker;
the redness increased by 17%; while the yellowness FTIR spectroscopy
reduced by 27% in comparison with the untreated
fiber. Furthermore, RPFs that were given H2SO4 The extent of chemical changes associated with the
treatment had brightness that is 6% lower compared to use of alkaline and acid treatments of RPFs was
those subjected to NaOH treatment. The redness of analyzed using FTIR spectroscopy. Typical absor-
acid treated fibers is 60% higher than that of alkaline bance spectra obtained from this study are presented in
treated fibers. However, the yellowness of acid treated Fig. 5 for untreated and treated fibers, with the
fibers is 15% lower than that of alkaline treated fibers. absorbance peaks of interest clearly marked. The
The change in colour of alkaline and acid treated RPFs peaks shown in Fig. 5 have been baseline-corrected
is due mainly to the removal of non-cellulosic and normalized to 2900 cm-1.
constituents of RPFs which are reported to be
Table 2 Variation of colour parameters of raffia palm fibers with NaOH and H2SO4 treatment
Fiber DL* Da* Db* DE
123
Cellulose
Fig. 3 Typical scanning electron microscope micrographs of the transverse section of a untreated, b alkaline treated and c acid treated
raffia palm fibers
Some of the characteristic peaks are summarized in gain in cellulose content after both chemical treat-
Table 3. In general, the spectra of the treated RPFs are ments as presented in Table 1. It is an indication of the
similar to that of the untreated fiber. However, after removal of some alkaline and acid-sensitive compo-
NaOH and H2SO4 treatments, there were reductions in nents of natural raffia palm fibers, thereby increasing
the intensity of certain characteristic peaks in com- the number of hydroxyl groups (Das and Chakraborty
parison with the untreated fiber. 2006). The –OH groups were reported to be involved
The absorption peak at approximately 3380 cm-1 in the formation of hydrogen bonds with carboxyl
in the untreated raffia palm fiber is attributed to O–H groups found on the surface of natural fibers
stretching vibrations in lignin (Kubo and Kadla 2005; (Mwaikambo and Ansell 2002; Hao 2013). The
Xu et al. 2013) and carbohydrate (cellulose and intensity of the absorption peak at approximately
hemicellulose) (Das and Chakraborty 2006; Chen 2900 cm-1 in the untreated fiber, which is attributed to
et al. 2009; Fahma et al. 2010; Ramadevi et al. 2012; C–H and CH2 stretching in hemicellulose, cellulose
Sathishkumar et al. 2014). It indicates the presence of and lignin (Himmelsbach and Akin 1998; Taha et al.
O–H groups in RPFs (Ouajai and Shanks 2005; Tibolla 2007; Tibolla et al. 2014) did not change significantly
et al. 2014). Its peak position shifted to the right to a with chemical treatment as that at 3380 cm-1. The
lower wavenumber (* 3337 cm-1) while its intensity absorption peak at approximately 1730 cm-1 in the
increased after NaOH and H2SO4 treatments. The untreated fiber is attributed to C=O stretching of
increase in intensity can be related to the observed methyl ester and carboxylic acid in pectin (Ouajai and
123
Cellulose
Fig. 4 Typical scanning electron microscope micrographs showing the surface of a untreated b alkaline treated and c acid treated raffia
palm fibers
Shanks 2005) or C=O stretching of acetyl and uronic the peak at * 1510 cm-1 in the untreated raffia palm
ester groups in hemicellulose (Cherian et al. 2008; fibers, attributed to C=C stretching of aromatic rings in
Chen et al. 2011; Xu et al. 2013). Its disappearance lignin (Moore and Owen 2001; Colom et al. 2003;
after NaOH treatment, which did not result in signif- Taha et al. 2007; Chen et al. 2011), remained after
icant reduction in hemicellulose content as shown in alkaline and acid treatments. This shows that none of
Table 1, suggests that it was pectin or wax and fats that the treatments was capable of removing lignin com-
was removed. Ouajai and Shanks (Ouajai and Shanks pletely from raffia palm fibers. Chemical treatment
2005) reported that pectin was removed from hemp caused a slight increase in the intensity of the peak
fibers after treatment in 8% NaOH solution at 30 °C at * 1425 cm-1, which is attributed to CH2 symmet-
for 1 h. The absorption peak at approximately ric bending in cellulose (Taha et al. 2007; Yue et al.
1606 cm-1 in the untreated raffia palm fiber is 2012; Xu et al. 2013). The intensity of the absorption
attributed to the adsorbed free water from cellulose peak at 1370 cm-1 in the untreated fiber, attributed to
in the fiber (Carrillo et al. 2004; Tibolla et al. 2014) aliphatic C–H bending in cellulose, hemicellulose and
and it also reveals aromatic ring vibration and C=O lignin (Kondo and Sawatari 1996; Carrillo et al. 2004;
stretching in lignin (Moore and Owen 2001; Kubo and Chen et al. 2010; Xu et al. 2013) increased in the acid
Kadla 2005; Cherian et al. 2008; Xu et al. 2013). Its treated fibers but was reduced in the alkaline treated
intensity decreased slightly with alkaline treatment but fibers. The intensity of the peak at * 1320 cm-1 in
increased slightly with acid treatment. The intensity of the untreated fiber, which is attributed to C–H and C–
123
Cellulose
123
Cellulose
Table 3 Main infrared transitions obtained for untreated and treated raffia palm fibers
Wave number (cm-1) Vibration assignment Source
123
Cellulose
Table 5 Analysis of variance results of the effect of treatment on tensile strength and percent elongation of raffia palm fibers
Property Source of variation df SS MS F p value
time and temperature used in the present study, both saturation occurred earlier than in the untreated fibers
NaOH and H2SO4 solutions were able to penetrate into between 106 and 178 h, while the onset of the second
the crystalline regions of the cellulose and degraded saturation occurred later than in the untreated fibers
the inter- and intra-hydrogen bonds between cellulose (between 298 and 494 h). For the acid treated fibers,
molecules and the adjoining crystalline regions (Chae the onset of the first saturation occurred earlier than for
et al. 2003; Suryanto et al. 2014). Reports about the untreated and alkaline treated fibers (between 82
degradation of tensile strength of natural fibers after and 154 h), while the second saturation occurred
chemical treatment abound in the open literature. between 274 and 494 h. The rapid water uptake
Symington et al. (Symington et al. 2009) reported that observed for the fibers at the early stage is attributed to
tensile strengths of flax and abaca fibers reduced with the presence of large lumens at the inner region of the
increasing treatment time in 3% NaOH solution. The fibers (see Fig. 3) which facilitated the rapid absorp-
tensile strains to failure for flax, abaca, kenaf and sisal tion of water by capillary action (Jacob et al. 2005).
fibers also decreased with increasing soaking time in The continued exposure of the fibers to water allowed
3% NaOH. Edeerozey et al. (Edeerozey et al. 2007) the internal pores of the elemental fibers to absorb
reported that increasing alkali concentration from 6 to water and this process occurred much more gradually
9% resulted in subtantial reduction in the tensile than at the initial stage of immersion. This probably
strength of kenaf fibers (from 239 to 165 MPa). Chae led to the second saturation stage. The average
et al. (Chae et al. 2003) reported a continual decrease percentage of water absorbed by the untreated fiber
in the break strength of lyocell fibers with increasing at the first and second saturation stages are 56 wt% and
concentration of alkaline. Similar results were 62 wt%, respectively.
reported for kenaf fibers by Mahjoub et al. (Mahjoub With chemical treatment, the average equilibrium
et al. 2014). They reported a decrease in the tensile water content for both first stage and second stage
strength of kenaf fiber due to increasing concentration saturation processes dropped. For alkaline treated
(5, 7, 10 and 15%) of NaOH solution and immersion fibers, it is 39 wt% for the first stage and 44 wt% for the
time. second stage. Since alkalization did not result in the
removal of either lignin or hemicellulose (see
Water uptake Table 1), it is believed that the observed reduction in
water content is due to the removal of pectin, wax and
Figure 7 shows the water uptake plots obtained for oily components of the fibers. On the other hand, the
untreated and chemically treated fibers. Both the percentage of water absorbed in the acid treated fibers
untreated and treated fibers experienced a two-stage at the first and second saturation stages are 30 wt% and
saturation behaviour during exposure. For the 35 wt%, respectively. Acid treatment removed hemi-
untreated fiber, first saturation occurred between 130 cellulose, pectin, wax and oils as shown in Table 1.
and 202 h and the second occurred between 274 and The further decrease in the maximum absorbed water
494 h. For the alkaline treated fibers, the first content (i.e., second saturation stage water content) in
123
Cellulose
X-ray diffraction
123
Cellulose
Table 6 Crystallite size and crystallinity index obtained for hydrolyze the amorphous fraction of cellulose con-
untreated and treated raffia palm fibers tained in NF, thereby leaving a higher fraction of
Material K k (Å) b (°) h (°) D (Å) CI (%) crystalline cellulose (Zafeiropoulos et al. 2002; Le
Troedec et al. 2008; Kaushik et al. 2012; Cai et al.
Cellulose 0.90 1.54056 4.00 22.9 21.56 59.1 2016).
Untreated 0.90 1.54056 2.83 22.2 30.32 54.4
fiber
Thermal analysis
NaOH treated 0.90 1.54056 1.42 22.4 60.51 68.2
H2SO4 0.90 1.54056 1.57 22.4 54.73 72.4
treated
Figures 9 shows the typical DSC curves obtained for
both the untreated and chemically treated RPFs, while
Table 7 summarizes the key reaction peak tempera-
material was determined using the Scherrer Eq. (3) tures. The untreated and chemically treated RPFs
(Nam et al. 2016). exhibit similar DSC profiles. For each fiber sample,
are three main peaks: one endothermic peak and two
Kk exothermic peaks. The broad endotherm is due to
D¼ ð3Þ
b cos h evaporation of water molecules in the fiber (Sara-
where D is the crystallite size, while K, k, b and h are vanakumar et al. 2013). Its enthalpy value and peak
the correction factor, wavelength, FWHM (in rad) and temperature changed with treatment. The two exother-
Bragg angle (i.e. 2h 7 2), respectively. The results are mic peaks are characteristics of natural fibers. Two
summarized in Table 6. The crystallite size calculated exothermic reactions at approximately 340 °C and
using the theoretical FWHM value of 4.0° is added for 475 °C were observed in the DSC thermogram of
comparison. It is seen that crystallite sizes of fibers Akamatsu wood (Tsujiyama and Miyamori 2000). The
treated with NaOH and H2SO4 are greater than that of transition occurring at 340 °C was attributed to
the untreated fiber. The increase in the crystallite size polysaccharides, while that at 475 °C was attributed
of the treated fibers was reported to be the reason for to lignin. In a DSC study of different types of wood
the sharpness of the peaks after chemical treatments biomass, three exothermic peaks were observed at
(Suryanto et al. 2014; Nam et al. 2016). about 280–310 °C, 355–365 °C, and 410–430 °C
The crystallinity index (CI) of the fibers was (Bryś et al. 2016), which were respectively attributed
estimated using Eq. (4) (Segal et al. 1959). to the decomposition of hemicellulose, cellulose, and
lignin. DSC analysis of pure hemicellulose, cellulose
It Ia and lignin showed peak decomposition temperature at
CI ð%Þ ¼ 100 ð4Þ
It approximately 275 °C, 355 °C and 365 °C, respec-
where It is the total intensity at (2 0 0) (located at tively (Yang et al. 2007). In the present study,
2h * 22.2°) and Ia is attributed to the amorphous exothermic reaction peak temperatures occurred at
fraction (i.e. the minimum intensity between (1 1 0) about 275 °C and 352 °C in the untreated, both of
and (2 0 0) peaks) which is located at 2h * 18.5° in which increased to higher values with chemical
the present study. CI values estimated for the tested treatment as shown in Table 6. On the basis of the
materials are shown in Table 6. mentioned previous DSC studies (Tsujiyama and
The apparent crystallinity index (CI) of untreated Miyamori 2000; Yang et al. 2007; Bryś et al. 2016),
RPF is estimated to be 54.4%, which is lower than CI the first exothermic peak is attributed to hemicellulose
values of 66.3% obtained for curaua fiber (Tomczak decomposition while the second peak is pyrolysis of
et al. 2007) and 64% reported for Raffia textilis fiber cellulose.
(Elenga et al. 2009). However, after alkaline and acid It is believed that the removal of non-cellulosic
treatments, the CI values obtained increased to 62.8% components and the resulting increase in cellulose
and 72.4%, respectively. The increase in the crys- content after chemical treatments made the observed
tallinity index of treated RPF can be attributed to the increase in the reaction peak temperatures possible.
observed increase in cellulose content as shown in Removal of non-cellulosic materials resulted in high
Table 1. Alkaline and acid treatments are known to degree of crystalline structural order which, in turn,
imparted high resistance to thermal degradation to the
123
Cellulose
Fig. 9 Typical differential scanning calorimetry thermograms obtained for a untreated b alkaline treated and c acid treated raffia palm
fibers
Table 7 Thermal properties of untreated and treated raffia palm fibers obtained from differential scanning calorimetry analysis
Sample Endothermic peak (°C) First exothermic peak (°C) Second exothermic peak (°C)
fibers (Ouajai and Shanks 2005). Hao (Hao 2013) alkaline and acid treated RPFs. The degradation
reported an increase in decomposition temperature of profile of untreated RPFs is characterized by three
cellulose from 357 to 367 °C for hemp fibers after distinct peaks, which is consistent with results
chemical treatment, while Aziz and Ansell (Aziz and reported for other natural fibers (Arias et al. 2013;
Ansell 2004) reported that alkalization treatment Elenga et al. 2013; Dong et al. 2014; Oliveira and
increased the thermal stability of both kenaf and hemp D’Almeida 2014). The first peak at * 66 °C in the
fibers. DTG curve (Fig. 10b) is due to evaporation of
Figure 10 shows TGA results (both weight loss in absorbed water and corresponds to roughly a mass
% and DTG in wt%/°C) obtained for untreated, loss of 5% (Fig. 10a) (Tomczak et al. 2007; Elenga
123
Cellulose
Conclusions
123
Cellulose
5. Degradation temperatures of hemicellulose and Bledzki AK, Gassan J (1999) Composites reinforced with cel-
cellulose were affected differently by chemical lulose based fibers. Prog Polym Sci 24:221–274
Bonarski JT, Olek W (2011) Application of the crystalline
treatment. volume fraction for characterizing the ultrastructural
organization of wood. Cellulose 18:223–235
Bruno IJ, Cole JC, Edgington PR, Kessler M, Macrae CF,
Acknowledgments The authors hereby acknowledge the
McCabe P, Pearson J, Taylor R (2002) New software for
technical assistance received from Mr. Jason Maley and Dr.
searching the Cambridge structural database and visualiz-
Jianfeng Zhu of the Saskatchewan Structural Sciences Centre
ing crystal structures. Acta Crystallogr Sect B 58:389–397
(SSSC), University of Saskatchewan, Saskatoon with respect to
Bryś A, Bryś J, Ostrowska-Lige˛za E, Kaleta A, Górnicki K,
FTIR and XRD data analysis.
Głowacki S, Koczoń P (2016) Wood biomass characteri-
zation by DSC or FT-IR spectroscopy. J Therm Anal
Compliance with ethical standards
Calorim 126:27–35
Cai M, Takagi H, Nakagaito AN, Li Y, Geoffrey INW (2016)
Conflict of interest The authors declare that they have no
Effect of alkali treatment on interfacial bonding in abaca
conflict of interest.
fiber-reinforced composites. Compos Part A Appl Sci
Manuf 90:589–597
Carrillo F, Colom X, Suñol JJ, Saurina J (2004) Structural FTIR
References analysis and thermal characterisation of lyocell and vis-
cose-type fibres. Eur Polym J 40:2229–2234
Chae DW, Choi KR, Kim BC, Oh YS (2003) Effect of cellulose
Alemdar A, Sain M (2008) Biocomposites from wheat straw
pulp type on the mercerizing behavior and physical prop-
nanofibers: morphology, thermal and mechanical proper-
erties of lyocell fibers. Text Res J 73:541–545
ties. Compos Sci Technol 68:557–565
Chen Y, Liu C, Chang PR, Cao X, Anderson DP (2009) Bio-
Ankom Technology (2000) Acid detergent fiber in feeds—filter
nanocomposites based on pea starch and cellulose nano-
bag technique. Method 5:6–7
whiskers hydrolyzed from pea hull fibre: effect of
Ankom Technology (2011) Neutral detergent fiber in feeds—
hydrolysis time. Carbohydr Polym 76:607–615
filter bag technique. Method 6:10–11
Chen H, Ferrari C, Angiuli M, Yao J, Raspi C, Bramanti E
Ankom Technology (2013) Determining acid detergent lignin in
(2010) Qualitative and quantitative analysis of wood
beakers. Method 8:11–12
samples by fourier transform infrared spectroscopy and
Arias A, Heuzey MC, Huneault MA (2013) Thermomechanical
multivariate analysis. Carbohydr Polym 82:772–778
and crystallization behavior of polylactide-based flax fiber
Chen W, Yu H, Liu Y, Hai Y, Zhang M, Chen P (2011) Isolation
biocomposites. Cellulose 20:439–452
and characterization of cellulose nanofibers from four plant
Arsène M, Okwo A, Bilba K, Soboyejo ABO, Soboyejo WO
cellulose fibers using a chemical–ultrasonic process. Cel-
(2007) Chemically and thermally treated vegetable fibers
lulose 18:433–442
for reinforcement of cement-based composites. Mater
Chen Z, Xu Y, Shivkumar S (2018) Microstructure and tensile
Manuf Process 22:214–227
properties of various varieties of rice husk. J Sci Food
ASTM D3822M-14 (2014) Standard test method for tensile
Agric 98:1061–1070
properties of single textile fibers. ASTM International USA
Cherian BM, Pothan LA, Nguyen-Chung T, Mennig G, Kot-
ASTM D-15 (2015) Standard practice for calculating yellow-
taisamy M, Thomas S (2008) A novel method for the
ness and whiteness indices from instrumentally measured
synthesis of cellulose nanofibril whiskers from banana
color coordinates. ASTM International USA
fibers and characterization. J Agric Food Chem
ASTM E1164-12 (2017) Standard practice for obtaining spec-
56:5617–5627
trometric data for object-color evaluation 1. ASTM Inter-
Chinga-carrasco G, Aslan M, Sørensen BF, Madsen B (2011)
national USA
Strength variability of single flax fibres. J Mater Sci
Aziz SH, Ansell MP (2004) The effect of alkalization and fibre
46:6344–6354
alignment on the mechanical and thermal properties of
Cicala G, Cristaldi G, Latteri A (2010) Composites based on
kenaf and hemp bast fibre composites: part 1—polyester
natural fibre fabrics. Woven Fabr Eng. https://doi.org/10.
resin matrix. Compos Sci Technol 64:1219–1230
5772/10465
Benyahia A, Merrouche A (2014) Effect of chemical surface
Colom X, Carrillo F, Nogués F, Garriga P (2003) Structural
modifications on the properties of alfa fiber–polyester
analysis of photodegraded wood by means of FTIR spec-
composites. Polym Plast Technol Eng 53:403–410
troscopy. Polym Degrad Stab 80:543–549
Bilba K, Ouensanga A (1996) Fourier transform infrared spec-
Das M, Chakraborty D (2006) Influence of alkali treatment on
troscopic study of thermal degradation of sugar cane
the fine structure and morphology of bamboo fibers. J Appl
bagasse. J Anal Appl Pyrolysis 38:61–73
Polym Sci 102:5050–5056
Bjerre AB, Olesen AB, Fernqvist T, Plöger AS, Schmidt AS
de Andrade Silva F, Chawla N, de Toledo Filho D (2008)
(1996) Pretreatment of wheat straw using combined wet
Tensile behavior of high performance natural (sisal) fibers.
oxidation and alkaline hydrolysis resulting in convertible
Compos Sci Technol 68:3438–3443
cellulose and hemicellulose. Biotechnol Bioeng
Dong Y, Ghataura A, Takagi H, Haroosh HJ, Nakagaito AN,
49:568–577
Lau KT (2014) Polylactic acid (PLA) biocomposites
reinforced with coir fibres: evaluation of mechanical
123
Cellulose
performance and multifunctional properties. Compos Part Kataoka Y, Kondo T (1998) FT-IR microscopic analysis of
A Appl Sci Manuf 63:76–84 changing cellulose crystalline structure during wood cell
Du R, Su R, Qi W, He Z (2018) Enhanced enzymatic hydrolysis wall formation. Macromolecules 31:760–764
of corncob by ultrasound-assisted soaking in aqueous Kaushik VK, Kumar A, Kalia S (2012) Effect of mercerization
ammonia pretreatment. Biotech 8:1–7 and benzoyl peroxide treatment on morphology, thermal
Edeerozey AMM, Akil HM, Azhar AB, Ariffin MIZ (2007) stability and crystallinity of sisal fibers. Int J Text Sci
Chemical modification of kenaf fibers. Mater Lett 1:101–105
61:2023–2025 Kondo T, Sawatari C (1996) A Fourier transform infra-red
Elenga RG, Dirras GF, Goma Maniongui J, Djemia P, Biget MP spectroscopic analysis of the character of hydrogen bonds
(2009) On the microstructure and physical properties of in amorphous cellulose. Polymer (Guildf) 37:393–399
untreated raffia textilis fiber. Compos Part A Appl Sci Kubo S, Kadla JF (2005) Hydrogen bonding in lignin: a fourier
Manuf 40:418–422 transform infrared model compound study. Biomacro-
Elenga RG, Djemia P, Tingaud D, Chauveau T, Maniongui JG, molecules 6:2815–2821
Dirras G (2013) Effects of alkali treatment on the Le Troedec M, Sedan D, Peyratout C, Bonnet JP, Smith A,
microstructure, composition, and properties of the raffia Guinebretiere R, Gloaguen V, Krausz P (2008) Influence of
textilis fiber. BioResources 8:2934–2949 various chemical treatments on the composition and
Fahma F, Iwamoto S, Hori N, Iwata T, Takemura A (2010) structure of hemp fibres. Compos Part A Appl Sci Manuf
Isolation, preparation, and characterization of nanofibers 39:514–522
from oil palm empty-fruit-bunch (OPEFB). Cellulose Lee CM, Kubicki JD, Fan B, Zhong L, Jarvis MC, Kim SH
17:977–985 (2015) Hydrogen-bonding network and OH stretch vibra-
Fiore V, Prestipino M, Valenza A, Scalici T, Nicoletti F, Vitale tion of cellulose: comparison of computational modeling
G (2015) A new eco-friendly chemical treatment of natural with polarized IR and SFG spectra. J Phys Chem
fibres: effect of sodium bicarbonate on properties of sisal 119:15138–15149
fibre and its epoxy composites. Compos Part B Eng Li Y, Li G, Zou Y, Zhou Q, Lian X (2014) Preparation and
85:150–160 characterization of cellulose nanofibers from partly
French AD (2014) Idealized powder diffraction patterns for mercerized cotton by mixed acid hydrolysis. Cellulose
cellulose polymorphs. Cellulose 21:885–896 21:301–309
French AD, Santiago Cintrón M (2013) Cellulose polymorphy, Mabrouk AB, Kaddami H, Boufi S, Erchiqui F, Dufresne A
crystallite size, and the Segal crystallinity index. Cellulose (2012) Cellulosic nanoparticles from alfa fibers (Stipa
20:583–588 tenacissima): extraction procedures and reinforcement
Gonzalez AV, Cervantes-Uc JM, Olayo R, Herrera-Franco PJ potential in polymer nanocomposites. Cellulose
(1999) Effect of fiber surface treatment on the fiber-matrix 19:843–853
bond strength of natural fiber reinforced composites. Macrae CF, Bruno IJ, Chisholm JA, Edgington PR, McCabe P,
Compos Part B Eng 30:309–320 Pidcock E, Rodriguez-Monge L, Taylor R, van de Streek J,
Hao A (2013) Mechanical and thermal properties of kenaf Wood PA (2008) Mercury CSD 2.0—new features for the
polypropylene nonwoven composites. Ph.D. Thesis; visualization and investigation of crystal structures. J Appl
University of Texas, Texas, USA, pp 19–30 Crystallogr 41:466–470
Hashim MY, Amin AM, Marwah FMO, Othman HM, Yunus Mahjoub R, Yatim JM, Mohd Sam AR, Hashemi SH (2014)
MRM, Huat CN (2017) The effect of alkali treatment under Tensile properties of kenaf fiber due to various conditions
various conditions on physical properties of kenaf fiber. of chemical fiber surface modifications. Constr Build
J Phys Conf Ser. https://doi.org/10.1088/1742-6596/914/1/ Mater 55:103–113
012030 Martins MA, Pessoa JDC, Gonçalves PS, Souza FI, Mattoso
Himmelsbach DS, Akin DE (1998) Near-infrared Fourier- LHC (2008) Thermal and mechanical properties of the acai
transform raman spectroscopy of flax (Linum usitatissi- fiber/natural rubber composites. J Mater Sci 43:6531–6538
mum L.) stems. J Agric Food Chem 46:991–998 Mohammed L, Ansari MNM, Pua G, Jawaid M, Islam MS
Holbery J, Houston D (2006) Natural-fibre-reinforced polymer (2015) A review on natural fiber reinforced polymer
composites in automotive applications. JOM 58:80–86 composite and its applications. Int J Polym Sci. https://doi.
Jacob M, Varughese KT, Thomas S (2005) Water sorption org/10.1155/2015/243947
studies of hybrid biofiber-reinforced natural rubber bio- Moore AK, Owen NL (2001) Infrared spectroscopic studies of
composites. Biomacromolecules 6:2969–2979 solid wood. Appl Spectrosc Rev 36:65–86
Jin W, Singh K, Zondlo J (2013) Pyrolysis kinetics of physical Morán JI, Alvarez VA, Cyras VP, Vázquez A (2008) Extraction
components of wood and wood-polymers using isocon- of cellulose and preparation of nanocellulose from sisal
version method. Agriculture 3:12–32 fibers. Cellulose 15:149–159
Jústiz-Smith NG, Virgo GJ, Buchanan VE (2008) Potential of Mukhopadhyay S, Fangueiro R, Shivankar V (2009) Variability
Jamaican banana, coconut coir and bagasse fibres as of tensile properties of fibers from pseudostem of banana
composite materials. Mater Charact 59:1273–1278 plant. Text Res J 79:387–393
Kabir MM, Wang H, Lau KT, Cardona F (2012) Chemical Mwaikambo LY, Ansell MP (2002) Chemical modification of
treatments on plant-based natural fibre reinforced polymer hemp, sisal, jute, and kapok fibers by alkalization. J Appl
composites: an overview. Compos Part B Eng Polym Sci 84:2222–2234
43:2883–2892 Nam S, French AD, Condon BD, Concha M (2016) Segal
crystallinity index revisited by the simulation of X-ray
123
Cellulose
diffraction patterns of cotton cellulose Ib and cellulose II. Saravanakumar SS, Kumaravel A, Nagarajan T, Sudhakar P,
Carbohydr Polym 135:1–9 Baskaran R (2013) Characterization of a novel natural
Nayak S, Mohanty JR (2018) Influence of chemical treatment on cellulosic fiber from Prosopis juliflora bark. Carbohydr
tensile strength, water absorption, surface morphology, and Polym 92:1928–1933
thermal analysis of areca sheath fibers. J Nat Fibers. https:// Sathishkumar TP, Navaneethakrishnan P, Shankar S, Rajasekar
doi.org/10.1080/15440478.2018.1430650 R (2014) Investigation of chemically treated randomly
Nishiyama Y, Langan P, Chanzy H (2002) Crystal structure and oriented sansevieria ehrenbergii fiber reinforced isoph-
hydrogen-bonding system in cellulose Ib from synchrotron thallic polyester composites. J Compos Mater
X-ray and neutron fiber diffraction. J Am Chem Soc 48:2961–2975
124:9074–9082 Sathitsuksanoh N, Zhu Z, Wi S, Percival Zhang YH (2011)
Odera RS, Onukwuli OD, Atuanya CU (2015) Characterization Cellulose solvent-based biomass pretreatment breaks
of the thermo-microstructural analysis of raffia palm fibers highly ordered hydrogen bonds in cellulose fibers of
proposed for roofing sheet production. J Miner Mater switchgrass. Biotechnol Bioeng 108:521–529
Charact Eng 3:335–343 Segal L, Creely JJ, Martin AE, Conrad CM (1959) An empirical
Oliveira AKF, D’Almeida JRM (2014) Characterization of method for estimating the degree of crystallinity of native
ubuçu (Manicaria saccifera) natural fiber mat. Polym cellulose using the X-ray diffractometer. Text Res J
Renew Resour 5:13–28 29:786–794
Ouajai S, Shanks RA (2005) Composition, structure and thermal Shibata S, Cao Y, Fukumoto I (2005) Effect of bagasse fiber on
degradation of hemp cellulose after chemical treatments. the flexural properties of biodegradable composites. Polym
Polym Degrad Stab 89:327–335 Compos 26:689–694
Panthapulakkal S, Zereshkian A, Sain M (2006) Preparation and Soleimani M, Tabil L, Panigrahi S, Opoku A (2008) The effect
characterization of wheat straw fibers for reinforcing of fiber pretreatment and compatibilizer on mechanical and
application in injection molded thermoplastic composites. physical properties of flax fiber-polypropylene composites.
Bioresour Technol 97:265–272 J Polym Environ 16:74–82
Persson J, Chanzy H, Sugiyama J (1991) Combined infrared and Soleimani M, Tabil LG, Oguocha I, Fung J (2017) Interactive
electron diffraction study of the polymorphism of native influence of biofiber composition and elastomer on phy-
celluloses. Macromolecules 24:2461–2466 sico-mechanical properties of PLA green composites.
Politou AS, Morterra C, Low MJD (1990) Infrared studies of J Polym Environ. https://doi.org/10.1007/s10924-017-
carbons Xiii. The oxidation of polycarbonate chars. Carbon 0967-8
N Y 28:855–865 Suryanto H, Marsyahyo E, Irawan YS, Soenoko R (2014) Effect
Ramadevi P, Sampathkumar D, Srinivasa CV, Bennehalli B of alkali treatment on crystalline structure of cellulose fiber
(2012) Effect of alkali treatment on water absorption of from mendong (Fimbristylis globulosa) straw. Key Eng
single cellulosic abaca fiber. BioResources 7:3515–3524 Mater 594–595:720–724
Reddy N, Yang Y (2005) Properties and potential applications Symington MC, Banks WM, West OD, Pethrick RA (2009)
of natural cellulose fibers from cornhusks. Green Chem Tensile testing of cellulose based natural fibers for struc-
7:190–195 tural composite applications. J Compos Mater
Reddy N, Yang Y (2007) Structure and properties of natural 43:1083–1108
cellulose fibers obtained from sorghum leaves and stems. Taha I, Steuernagel L, Ziegmann G (2007) Optimization of the
J Agric Food Chem 55:5569–5574 alkali treatment process of date palm fibres for polymeric
Rokbi M, Osmani H, Imad A, Benseddiq N (2011) Effect of composites. Compos Interfaces 14:669–684
chemical treatment on flexure properties of natural fiber- Thiruchitrambalam M, Athijayamani A, Sathiyamurthy S,
reinforced polyester composite. Procedia Eng Thaheer ASA (2010) A review on the natural fiber-rein-
10:2092–2097 forced polymer composites for the development of roselle
Ronald Aseer J, Sankaranarayanasamy K, Jayabalan P, fiber-reinforced polyester composite. J Nat Fibers
Natarajan R, Priya Dasan K (2013) Morphological, phys- 7:307–323
ical, and thermal properties of chemically treated banana Tibolla H, Pelissari FM, Menegalli FC (2014) Cellulose nano-
fiber. J Nat Fibers 10:365–380 fibers produced from banana peel by chemical and enzy-
Rosa MF, Chiou B, Medeiros ES, Wood DF, Williams TG, matic treatment. LWT Food Sci Technol 59:1311–1318
Mattoso LHC, Orts WJ, Imam SH (2009) Effect of fiber Tomczak F, Satyanarayana KG, Sydenstricker THD (2007)
treatments on tensile and thermal properties of starch/ Studies on lignocellulosic fibers of Brazil: part III—mor-
ethylene vinyl alcohol copolymers/coir biocomposites. phology and properties of Brazilian curaua fibers. Compos
Bioresour Technol 100:5196–5202 Part A Appl Sci Manuf 38:2227–2236
Sampathkumar D, Punyamurth R, Venkateshappa SC (2012) Tsujiyama SI, Miyamori A (2000) Assignment of DSC ther-
Effect of chemical treatment on water absorption of areca mograms of wood and its components. Thermochim Acta
fiber. J Appl Sci Res 8:5298–5305 351:177–181
Sandy M, Bacon L (2001) Tensile testing of raffia. J Mater Sci Vardhini KJV, Murugan R, Selvi CT, Surjit R (2016) Optimi-
Lett 20:529–530 sation of alkali treatment of banana fibres on lignin
Santhosh J, Balanarasimman N, Chandrasekar R, Raja S (2014) removal. Indian J Fibre Text Res 41:156–160
Study of properties of banana fiber reinforced composites. Xu F, Yu J, Tesso T, Dowell F, Wang D (2013) Qualitative and
Int J Res Eng Technol 3:144–150 quantitative analysis of lignocellulosic biomass using
123
Cellulose
infrared techniques: a mini-review. Appl Energy from native and mercerized cotton fibers. Cellulose
104:801–809 19:1173–1187
Yan L, Chouw N, Yuan X (2012) Improving the mechanical Zafeiropoulos NE, Williams DR, Baillie CA, Matthews FL
properties of natural fibre fabric reinforced epoxy com- (2002) Engineering and characterisation of the interface in
posites by alkali treatment. J Reinf Plast Compos flax fibre/polypropylene composite materials. Part I.
31:425–437 Development and investigation of surface treatments.
Yang H, Yan R, Chen H, Lee DH, Zheng C (2007) Character- Compos Part A Appl Sci Manuf 33:1083–1093
istics of hemicellulose, cellulose and lignin pyrolysis. Fuel
86:1781–1788
Publisher’s Note Springer Nature remains neutral with
Yue Y, Zhou C, French AD, Xia G, Han G, Wang Q, Wu Q
regard to jurisdictional claims in published maps and
(2012) Comparative properties of cellulose nano-crystals
institutional affiliations.
123