You are on page 1of 15

Appl Compos Mater (2007) 14:89103

DOI 10.1007/s10443-006-9032-9

Study of Structural Morphology of Hemp Fiber


from the Micro to the Nanoscale
Bei Wang & Mohini Sain & Kristiina Oksman

Received: 9 October 2006 / Accepted: 27 December 2006 /


Published online: 30 January 2007
# Springer Science + Business Media B.V. 2007

Abstract The focus of this work has been to study how high pressure defibrillation and
chemical purification affect the hemp fiber morphology from micro to nanoscale.
Microscopy techniques, chemical analysis and X-ray diffraction were used to study the
structure and properties of the prepared micro and nanofibers. Microscopy studies showed
that the used individualization processes lead to a unique morphology of interconnected
web-like structure of hemp fibers. The nanofibers are bundles of cellulose fibers of widths
ranging between 30 and 100 nm and estimated lengths of several micrometers. The chemical
analysis showed that selective chemical treatments increased the -cellulose content of
hemp nanofibers from 75 to 94%. Fourier transform infrared spectroscopy (FTIR) study
showed that the pectins were partially removed during the individualization treatments.
X-ray analysis showed that the relative crystallinity of the studied fibers increased after
each stage of chemical and mechanical treatments. It was also observed that the hemp
nanofibers had an increased crystallinity of 71 from 57% of untreated hemp fibers.
Key words cellulose nanofibers . hemp . microfibrils . nanostructures . characterization

1 Introduction
Lately, there has been considerable interests in the isolation and study of novel
nanomaterials manufactured from renewable resources. An important class of nanomaterials
has been nanofibers and fibrils from different cellulose sources and cellulose crystals
B. Wang : M. Sain (*) : K. Oksman
Centre for Biocomposites and Biomaterials Processing, Faculty of Forestry,
University of Toronto Earth and Science Centre, 33 Willcocks Street, Toronto, ON M5S 3B3, Canada
e-mail: m.sain@utoronto.ca
B. Wang
e-mail: bei.wang@utoronto.ca
K. Oksman
Manufacturing and Design of Wood and Bionanocomposites, Lule University of Technology,
SE-93187 Skellefte, Sweden

DO9032; No of Pages

90

Appl Compos Mater (2007) 14:89103

(whiskers) [14]. These novel nanofibers, fibrils and crystals have been shown to result in
unique properties when incorporated in different polymers [1, 58]. The sources of these
nanomaterials have been wheat straw, bacterial cellulose, kraft pulp, sugar beet pulp, potato,
and swede root [2, 5, 710]. Since 1998, Canada has grown industrial hemp for seed and
for fiber. Interest in hemp arises from the plants amazing versatility. The seeds are used to
produce healthy food, nutraceuticals, and bodycare products and the stalk is starting to be
processed into high performance fiber products such as paper, textiles, biocomposites and
building materials [11].
The chemical constituents of the hemp plants cell wall consist not only of cellulose, but
also of hemicellulose, pectin and lignin. The properties of each constituent contribute to the
overall properties of the fiber [12]. The smallest building element of the cellulose skeleton
is considered by some to be an elementary fibril. The fibril can be about 510 nm in
diameter and its length varies from 100 nm to several micrometers depending of the source
of cellulose [13]. The cellulose molecules are always biosynthesized in the form of
nanosized fibrils; up to 100 glucan chains aggregate together to form cellulose nano-sized
microfibrils or nanofibers [1417]. The mechanical performance of cellulose nanofibers in
terms of the tensile strength and Youngs modulus is comparable to other engineering
materials such as glass fiber, carbon fiber, etc. Therefore, the cellulose nanofibers can be
considered to be an important structural element of natural cellulose in a number of
applications such as plastic reinforcement, gel forming and thickening agents [3, 18, 19].
Furthermore, a cellulose nanofiber has more than 200 times the surface area of isolated
softwood cellulose [20] and possesses higher water holding capacity, higher crystallinity,
higher tensile strength, and a finer web-like network. In combination with a suitable matrix
polymer, cellulose nanofiber networks show considerable potential as an effective
reinforcement for high quality specialty application of bio-based composites. Another type
of nanoreinforcement that can be obtained from cellulose fibers are nanowhiskers. The
elementary fibril is made up of amorphous and crystalline parts. The crystalline parts can be
isolated by various treatments producing the cellulose nanowhiskers [4, 21].
Many studies have been done on extracting cellulose microfibrils from various sources and
on using them as reinforcement in composite manufacturing [1, 2, 6, 7, 2224]. These
microfibrils can be extracted from the cell walls by three types of isolation processes: simple
mechanical methods, a combination of chemical and mechanical methods, or an enzymatic
approach. A purely mechanical process can produce refined, fine fibrils several micrometers
long and between 20 to 90 nm in diameter [25]; however, this nano-scalar web-like structure
of fibrils causes a reduction of strength [5]. In contrast, chemi-mechanical treatments can
extract cellulose nanofibers from the primary and secondary cell walls without degrading the
cellulose. A chemi-mechanical process can also achieve finer fibrils of cellulose, ranging
between 5 and 60 nm diameter [1, 7]. Depending upon the raw materials and defibrillation
techniques, the degree of polymerization, morphology and aspect ratio of the nanofibers will
differ. The separation of nanoreinforcement from natural materials and the processing
techniques have been limited to laboratory scale [13]. Therefore, it is important to develop
new processing techniques which will be at use in large scale production.
Removal of lignin left after chemical treatment of fibers is the goal of the bleaching
process. Chlorine-based processes still dominate, but more environmentally friendly nonchlorine processes are becoming more prevalent. A bleaching treatment using a sodium
chlorite solution was performed to remove phenolic compounds or molecules having
chromophore groups, in order to whiten the fibers [1]. This is a popular technique at the
laboratory scale to remove lignin from plants. Lignin is rapidly oxidized by chlorine and
chlorites. Lignin oxidizing leads to lignin degradation and to dissolution in an alkaline

Appl Compos Mater (2007) 14:89103

91

medium [1]. A further treatment is often required to fully bleach the suspension. Removal
of a part of the flax fibers noncellulosic compounds by sodium chlorite was reflected in the
mechanical and physical characteristics of the surface state [26].
Key issues are the size and the dispersion of the nano-sized reinforcements and the effect
of this fine structure on fiber properties. Transmission electron microscopy (TEM) and
atomic force microscopy (AFM) aid the interpretation of structures from the nm to the m
size scale. Typical information obtained from conventional TEM is length, aspect ratio,
shape and the aggregated or isolated state of fibers [27]. Thin evaporative carbon coatings
were used for TEM sample preparation in this study. Recently, AFM has also been used to
examine plant cell walls at a similar resolution to that of the TEM [2830]. This type of
microscopy has the important advantage of reducing the risks of introducing artefacts
resulting from the preparative techniques.
In order to investigate the potential of cellulose nanofiber as a reinforcement in polymer
composites, this study was focused on the development of a new isolation technique to
extract cellulose nanofibers from hemp by chemi-mechanical treatments. This research aims
to clarify how the various levels of high pressure defibrillation affect the morphology from
long hemp fiber towards nano-scale fibrillated cellulose and to compare the morphology of
bleached and unbleached fibers at the different stage of the individualization of the
nanofibers. The structural details were studied with SEM, TEM and AFM. The crystallinity
was determined before and after different stages of the chemi-mechanical treatments of
hemp fibers. The changes of the chemical composition of fibers after different treatments
were studied. Infrared measurements were performed to identify the removal of the pectins.
Further research work is required for incorporating these nanofibers into a polymer matrix
to evaluate the mechanical properties of nanocomposites.

2 Experimental
2.1 Materials and Methods
The raw material used in this study was hemp fibers (Cannabis sativa L.) from southwestern
Ontario, Canada (Hempline, Ontario). These fibers have diameters of approximately 2225 m
and lengths of 1525 mm.
Reagent grade chemicals were used for fiber surface modifications and bleaching,
namely, sodium hydroxide, hydrochloric acid, sodium chlorite, chlorine dioxide, peroxide
and sulfuric acid.
2.2 Individualization Process
The individualization process of nanofibers is a multi-step process, shown in Fig. 1.
Chemical and mechanical treatments together are applied onto the hemp fibers to
individualize nanofibers. The chemical treatments include pre-treatment, acid hydrolysis,
alkaline treatment and bleaching. The mechanical treatments include cryo-crushing by
liquid nitrogen and high-pressure defibrillation [1].
2.2.1 Chemical Treatments
The main objective of the chemical treatments was to remove the starch, hemi-cellulose,
lignin/pectins surrounding cellulose. Generally, the first step for all of the fiber surface

92

Appl Compos Mater (2007) 14:89103

Raw Material (hemp)


Successive Bleaching
Pretreatment (12% w/w NaOH, 2h)
Chlorine Dioxide Stage
Acid Hydrolysis (1M HCl, 80C, 1.5h)
Extraction Stage
Alkaline Treatment (2 % w/w NaOH, 2h, 80C)
Acid Stage

Cryo-crushing in Liquid Nitrogen


Peroxide Stage

High Pressure Defibrillation


Fig. 1 Isolation of nanofibers

treatments is mercerization (pretreatment process) which will change the crystal structure of
cellulose. The essence of mercerizing hemp is that in the swelling of cellulose fibers due to
exposure to alkalis, the natural crystalline structure of the cellulose relaxes and under an
appropriate tension, the dimensions can be set by the conditions [31].
Hemp fibers were soaked in a sodium hydroxide solution of 12% w/w at room
temperature for 2 h, enabling chemical molecules to penetrate through the crystalline
region of the cellulose. Acid hydrolysis with 1 M hydrochloric acid followed by alkaline
treatment with 2% w/w sodium hydroxide was applied to remove the undesired components.
After the successive chemical treatments, lignin was still remained within the fibers and
removed by multi-stage bleaching. The bleaching was done in four different stages:
1) Chlorine dioxide stage (D), where the fiber consistency was adjusted to 3.5%. Sodium
chlorite solution was applied based on the Kappa number of the fibers. Therefore the
Kappa number was determined and then chlorine dioxide was calculated based on
lignin content in the sample. The retention time was 1 h.
2) Extraction stage (E), where the consistency was adjusted to 10% using boiling water.
Sodium hydroxide and peroxide was added to the fiber stock based on 1% OD fiber
under mechanical stirring.
3) Acid stage (A), where the consistency was adjusted to 4% and sulfuric acid was added
to fiber mixing well for 1 h.
4) Peroxide stage (P), where the consistency was adjusted to 10% and peroxide was
added to fiber. After filtration, the fiber was washed and air dried.

Appl Compos Mater (2007) 14:89103

93

Process for chemical pulping was obtained from literature [32]. However, at this
purification level, microfibrils are not individualized and further steps of mechanical
treatments are needed.
2.2.2 Mechanical Treatments
The hemp fibers were first cryo-crushed with liquid nitrogen to reduce the length and size. The
objective of the cryo-crushing is to form ice crystals within the fiber cell wall. When high
impact is applied on the frozen fiber, ice crystals exert pressure on the cell wall, causing a
rupture and thereby liberating the microfibrils [33]. The chemically treated and cryo-crushed
fibers were then diluted in water and dispersed evenly in a disintegrator (Cramer) for 10 min.
The disintegrator is used to disperse the fibers uniformly in the water suspension before the
high pressure defibrillation process. The water suspension with higher concentration of fibers
(12%) was subsequently passed through the defibrillator (purchased from GEA-modified by
the Centre for Biocomposites and Biomaterials Processing Laboratory, University of Toronto,
Ontario). The pressure was above 500 bar and several passes were needed to crush the cell
wall and fully release the nanofibers. The detailed method for mechanical treatment to
produce the nanofibers is described in recent patents [33, 34].
2.3 Microscopy Characterization
Scanning electron microscope (JEOL JSM-840, Tokyo, Japan) was used as a routine for
microstructural analysis of the fibers after various stages of chemical and mechanical
treatments. All images were taken at an accelerating voltage of 15 kV. The sample surfaces
were coated with a thin layer of gold on the surface using an Edwards S150B sputter coater
(BOC Edwards, Wilmington, MA) to provide electrical conductivity.
Transmission electron microscopy (TEM) observations were achieved with a Philips
CM201 (Philips, Eindhoven, The Netherlands) operated at 80 kV. A drop of a dilute
cellulose nanofiber suspension was deposited on carbon-coated grids and allowed to dry.
Atomic force microscope (AFM) study was obtained using a digital instruments
dimension 3100 AFM (Veeco Metrology Group, Santa Barbara, CA) with a nanoscope IIIa
controller. The system was operated in tapping mode at room temperature with DI tapping
mode tips having a resonant frequency of 280 kHz. A droplet of the aqueous nanofiber
suspension was allowed to dry on a cleaved mica surface.
2.4 Chemical Characterization of Fibers
Over the different stages of nanofiber development, untreated hemp fibers, acid/alkali
treated fibers, bleached fibers and nanofibers were chemically analyzed for hemicellulose,
lignin and cellulose contents. The procedure used here for cellulose determination was
given by Zobel et al. [35]. Lignin content was determined based on Tappi T 222 om02.2002 (acid-insoluble lignin in wood and pulp) and Tappi useful method UM250 (raw
material and pulp-determination of acid-soluble lignin).
2.5 Spectroscopy
Fourier transform infrared spectroscopy (FTIR), Tensor 27, Bruker Optics, Billerica,
MA) was used to identify the removal of pectins at different purification levels by
measuring the transmitted radiation of various infrared light wavelengths of pectin

94

Appl Compos Mater (2007) 14:89103

functional groups in the sample. The Tensor 27 standard FTIR spectroscopy was used to
obtain spectra for the fibers after each chemical treatment. Fibers were ground and mixed
with KBr (sample/KBr ratio, 1/99) to prepare pastilles. FTIR spectra were recorded in a
spectral range of 4,000400 cm1 with a resolution of 4 cm1.
2.6 X-Ray Analysis
The crystallinity determination was made using a powder X-ray diffraction method
(PXRD). X-ray crystallography was carried out to investigate the relative crystallinity after
various stages of the chemi-mechanical treatment and of nanofibers obtained by air drying
of the nanofiber suspension. A D8 advance Bruker AXS diffractometer (Bruker AXS,
Madison, WI), Cu point focus source, graphite monochromator, and 2D-area detector
GADDS system were used. Samples were analyzed in transmission mode.

3 Results and Discussion


3.1 Individualization of Hemp Nanofibers
The individualization step of nano-sized fibers from the plant cell walls requires chemical
and mechanical treatments. The properties of the nanomaterials cannot be physically
measured without separating them from the plant cell wall. These treatments may result in
significant chemical or mechanical damage on the fibers.
Figure 2 shows how the fiber morphology is changed from the micro to the nanoscale
during the individualization process. Figure 2a shows an untreated hemp fiber bundle where
the individual fibers are bound together by lignin. The size of the bundle is around 100
200 m. In Fig. 2b, it is clearly visible that the chemical treatments are reducing the bundle
size and the surface roughness compared to the fibers in Fig. 2a. Figure 2c shows how the
morphology is affected by the cryo-crushing. This process imparted sufficient energy to
break the bundles into single fibers which are around 20 m in width. In Fig. 2d, the single
fibers are defibrillated showing a web-like structure. No individual hemp fibers are visible
after the defibrillation step and the size is reduced to the nanometer level. The fibrillar
structure of individual fibers was revealed from the Fig. 2e after the fiber bleaching and
may be due to the leaching out of waxes and pectic substances. In Fig. 2f, it was observed
that the diameter of cryo-crushed fibril with bleaching is much smaller compared to cryocrushed fibril without bleaching. High pressure defibrillation provided high turbulence and
shear that created an efficient mechanism of reduction in size. Figure 2g shows the structure
after the high pressure defibrillation, showing nanoscale fibrils and microfibril bundles
contributed a unique morphology of the interconnected web-like structure of fibrils. This
combination of forces promoted a high degree of microfibrillation of cellulose fibers,
resulting in cellulose nanofibers.
Figure 3 demonstrates how the number of passes through a defibrillator is affecting the
individualization of hemp nanofibers. Figure 3a shows the morphology after five passes
which did not result in nanofiber structure. The fibers were still entangled with each other
and the size was in the range of microns. In Fig. 3b, after 10 passes, the fibers were split
apart into smaller bundles. Figure 3c and d show a large extent of defibrillation after 15 and
20 passes, these small bundles were additionally separated into thinner fibril bundles
increasing the exposed surface area of the cellulose (Fig. 3d). High pressure and high
energy were needed to defibrillate hemp fibers and achieve acceptable dispersion level. The

Appl Compos Mater (2007) 14:89103

95

Fig. 2 Scanning electron micrographs of: a untreated fiber, b after acid and alkaline treatment, c after cryocrushing, d after defibrillation, e after bleaching, f bleaching followed by cryo-crushing, g bleaching
followed by cryo-crushing and defibrillation

separated fiber bundles were shown to create small entanglements that were fibrillated into
smaller entities as the number of passes through the defibrillator was increased [5].
Figure 4, compares the structure of unbleached and bleached nanofibers. Figure 4a
shows that the unbleached nanofibers were either individual or bundles and coarser
compared to the bleached nanofibers, shown in Fig. 4b. The Fig. 4b shows that bleached
nanofibers were thinner and shorter than unbleached nanofibers in diameter and length. It is
likely that harsh chemicals used for bleaching will reduce the chain length of the cellulose
which can result in cutting the length and weakening cellulose nanofibers. Aspect ratio
(fibril length to diameter ratio) is one of the most important parameters in determining
reinforcing capability of the nanofibers. Aspect ratios of the extracted cellulose nanofibers
were estimated from transmission electron micrographs. In some cases total fibril length
was not visible, therefore only the visible portion was considered for the calculation in
TEM graphs (provide statistical significance of this assumption). The aspect ratio of these
bleached nanofibers (82) is comparable to unbleached nanofibers (88). We can expect a
high reinforcing capability from both nanofibers. There is a direct relation between degree
of polymerization and length of the nanofibers, as cellulose synthesizes in extended chain
conformation. Degree of polymerization (DP) of nanofibers was calculated using intrinsic
viscosity method (ASTM D179596). Unbleached nanofiber has a similar valve of DP
(1,155) compared with that of bleached nanofibers (1,138).
The nanofiber suspension obtained after the high-pressure defibrillation was also
analyzed to determine the width using AFM, shown in Fig. 5. It is seen that the fibers are
indeed nano-sized and the width is within the range of 30100 nm. The length is estimated
to be at micrometer level. The network of nanofibers can also be seen in force mode image

96

Appl Compos Mater (2007) 14:89103

Fig. 3 Scanning electron micrographs of: bleached hemp nanofibers after 5 (a), 10 (b), 15 (c) and 20 (d)
passes during the defibrillation

Fig. 5a by interwoven microfibrils overlapping each other. As evident from the TEM and
AFM images, high-pressure defibrillation leads to individualization of the cellulose
nanofibers from the cell wall without degrading them. Figure 6 shows the width
distribution of the nanofibers obtained through chemi-mechanical treatments. In Fig. 6a,
it was observed that the widths of unbleached nanofibers were estimated between 50
100 nm and most of them had a diameter range of 70 to 100 nm. Bleached nanofibers

Fig. 4 Transmission electron micrographs of hemp nanofibers (a) unbleached, (b) bleached under the same
magnification (15,000)

Appl Compos Mater (2007) 14:89103

97

Fig. 5 Atomic force micrographs of unbleached hemp nanofibers a force mode image, b height mode image

(a)
Frequencyof Width

0.4
0.3
0.2
0.1
0
Width Range (nm)

(b)
Frequencyof Width

Fig. 6 Size distribution of hemp


nanofibers a unbleached,
b bleached

<30

30-50

<30

30-50

50-70

70-100

>100

0.5
0.4
0.3
0.2
0.1
0
Width Range (nm)
50-70

70-100

>100

98

Appl Compos Mater (2007) 14:89103

Table 1 Chemical analysis of hemp fibers after selective chemical treatments

Untreated fibers
Acid treated
Acid and alkaline
treated
Bleached
Nanohemp

Holocellulose
(%)

-cellulose
(%)

Hemicellulose
(%)

Insoluble
lignin (%)

Soluble
lignin (%)

Total lignin
(%)

86.22
91.42
92.82

75.56
85.66
89.78

10.66
5.76
3.04

4.89
4.49
4.40

1.72
0.66
0.53

6.61
5.15
4.93

95.72
96.12

93.87
94.53

1.85
1.59

2.83
2.53

0.35
0.18

3.18
2.71

produced smaller widths (30100 nm) compared with that of unbleached nanofibers as
shown in Fig. 6b. Most of bleached nanofibers had a diameter range of 3050 nm. By the
fiber image statistical analysis, the average width of unbleached nanofiber was 87.5 nm
and of bleached nanofiber was 54 nm based on averaging 50 individual fibers. It is
expected that better reinforcing effect will be obtained by using bleached nanofibers. It is
clearly shown from The TEM picture (Fig. 4) and the fiber size distribution chart (Fig. 6)
that the size is less in bleached nanofibers, but more surface areas will be contributed from
these fibrils.
3.2 Chemical Characterization of Individualized Nanofibers
The fibers obtained after the chemical treatment contained mainly alpha-cellulose with
some hemi-cellulose and lignin. As shown in Table 1, the -cellulose content in the
chemically treated fibers was 94% as compared to the original 75% and the hemicellulose
content was reduced to 1.6%. Chemical analysis of these fibers after each stage of the
purification showed a drastic increase in cellulose content and a decline in hemicellulose
and lignin content. Lignocellulosic fibers contain a bit amount of hemicellulose, which is a
hetero polysaccharide consisting mainly of pentoses and hexoses. The treatment of
cellulosic, starch, or hemicellulosic materials using acid solution to break down the
polysaccharides to simple sugars allows the solubilization of both pectins and hemicelluloses. Dilute sodium hydroxide treatment of lignocellulosic fibers causes separation of
structural linkages between lignin and carbohydrate and disruption of lignin structure [32].
Fibers samples from hemp appeared brownish in color even after carrying out the acid and
alkali treatments. Chemically treated hemp fibers were bleached before proceeding to the
mechanical treatment to ensure that most of the lignin was removed from the fibers. After
bleaching treatment, it was found that nanofibers contain both soluble and insoluble lignin.
The lignin content of hemp significantly decreased from 6.6 to 3.18%. According to the
result of successive bleaching extractions, the nanofibers lost most of their non-cellulosic
constituents. Cellulose can be also partially degraded during these bleaching.
Chemical treatments lead to almost pure cellulose fibers, which ensure the high stiffness
and strength. The mechanical behavior of nanohemp reinforced composites changes as a
function of hemp cellulose microfibrils purity level [5]. Although cellulose possesses
excellent strength and good stability, it can be partially degraded due to the harsh chemical
and mechanical treatment. The alkali extraction is expected to hydrolyze pectin by a elimination process and solubilize it [1, 7]. Figure 7 shows the transmission electron
micrographs of hemp nanofibers under two controlled concentrations of NaOH solution for
the fiber pretreatment. 17.5% of sodium hydroxide induced undesirable reactions to cut
down the cellulose chains, therefore reducing the aspect ratio of nanofibers. Too harsh

Appl Compos Mater (2007) 14:89103

99

Fig. 7 Transmission electron micrographs of hemp nanofibers a under 17.5% alkali extraction, b under 12%
alkali extraction (15,000)

Transmittance (%)

treatments led to the loss of the microfibrillar morphology as seen in Fig. 7a. In Fig. 7b, it
was shown that the hemp fibers released their nanofibers either individually or in bundles at
a lower alkali content leading to the formation of a strong network of microfibrils.
Pectin is a heterogeneous grouping of acidic structural polysaccharides. The untreated,
chemically treated and bleached hemp fibers were characterized by FTIR, shown in Fig. 8.
By this technique, it was possible to follow the removal of pectins due to the vanishing of

1739
1259
895
1631

400

600

800

1000

1200

1400

1600

1800

2000

2200

Wavenumber cm-1
Untreated

Chemically treated

Bleached

Fig. 8 FTIR spectra of hemp fibers a untreated, b after acid and alkaline treatment and c after bleaching

100

Appl Compos Mater (2007) 14:89103

the characteristic bands for the carboxylate groups at 1,740 cm1, acetyl and methyl ester
group at 1,590 and 1,240 cm1 [7]. In the transmittance spectra, the absorption band
assigned to the pectin carboxylic groups was observed at 1,739 cm1 in untreated fibers, but
disappeared upon chemical treatments and the successive bleaching. This is because the
carboxylic groups were partially removed by alkali treatment through a process called deesterification. During de-esterification, the ester groups on the pectin can be removed as
well. The alkali treatment allows the ionization of pectin carboxylic groups (COOH) and
the formation of the corresponding sodium carboxylate (COONa), which decreases the
ability of hydrogen-type intermolecular bonds to form [7] and the solubility of the pectins.
Reduction in the peak intensity found at around 1,6311,633 cm1 in chemically treated
and bleached fibers indicates the partial reaction of the C=O bonds of hemicelluloses. The
intensity of the 1,259 cm1 peak is sharply weakened after the bleaching treatment, due to
the removal of hemicellulose materials. The peak observed at 895 cm1 in both untreated
and chemically treated fibers indicates the presence of the glycosidic linkages between the
monosaccharides and disappears in the spectrum for bleached fibers.
3.3 X-Ray Diffraction

Intensity (a.u.)

X-ray crystallography was used to investigate the crystallinity of the sample after different
treatments. X-ray powder diffraction photographs from untreated, acid and alkali treated
fibers and nanofibers are shown in Fig. 9. The percentage crystallinity of these samples was
calculated based on X-ray analysis by Eq. 1 and they are given in Fig. 9 as well. The main
diffraction intensity was at about 2=21 for each sample. The peak observed close to 2=
22.4 is from cellulose. Untreated fiber exhibited very low crystallinity (57.4%) and a

10

15

20

25

30

35

Diffraction angle 2 (degree)


Untreated (57.4%)
Acid & Alkaline treated (69.7%)

Acid treated (61.9%)


Nanofiber (71.2%)

Fig. 9 X-ray diffractometry and crystallinity estimation after each stage of chemo-mechanical treatment for
unbleached hemp fiber

Appl Compos Mater (2007) 14:89103

101

single peak at about 2=22.75 and a broad hump showing amorphous nature. It can be
seen that acid treated and acid/alkali treated fibers show peaks at 2=22.8 and 2=21.4
respectively. In the case of acid treated fibers, an additional peak was seen at 2=20.4 with
a small hump. It is possible that acid treated fiber contains some residual lignin and
hemicellulose that contribute to the slightly lower crystallinity (61.9%). Further the hump
has disappeared showing that amorphous chains have been rearranged to crystalline
regions. Acid/alkali treated fiber shows only one prominent peak and exhibits more
crystalline nature (69.7%) than acid treated fiber. Nanofibers showed a high crystalline
feature with a narrow strong peak at 2=21.6. This peak can be attributed to cellulose
crystals. It was observed that the hemp nanofibers had an increased crystallinity from
57.4% of untreated hemp fibers to 71.2% of nanofibers. As shown in Fig. 9, the relative
crystallinity of the samples increased after each stage of chemical treatments. Consecutive
chemical treatments and processing of cellulose fibers give different X-ray patterns.
According to X-ray testing on cellulose, cellulose is not made up of single perfect crystals.
Disordered cellulose molecules as well as hemicelluloses and lignin are located in the
spaces between the microfibrils. The hemicelluloses are considered to be amorphous
although they apparently are oriented in the same direction as the cellulose microfibrils.
Lignin is both amorphous and isotropic. It is believed that these crystallites are connected to
each other by disoriented amorphous zones [36]. The crystalline nature of the cellulose
nanofibers is not only influenced by the chain conformation but also by the packing of
adjacent chains. Nanofibers are pure cellulose chains having different arrangements of the
glucose chains as in the native cellulose of untreated fibers. The hemp nanofiber, after
successive chemical treatments, only possessed 4.3% lignin and hemicellulose combined.
The X-ray powder diffraction pattern showed a narrow peak which was more prominent
and sharp for nanofiber, indicating the crystalline nature of this reinforcement and higher
relative crystallinity.
C; %

Icrystalline
 100
Icrystalline Iamorphous

4 Conclusions
This study has been concerned how the degree of individualization affects the hemp fiber
morphology from the micro to the nanoscale. The chemi-mechanical process resulted in
hemp nanofibers having a width in the range of 30100 nm. The used chemical treatments
resulted in the individualized hemp microfibers and further mechanical treatment formed a
network structure of hemp nanofibers. The high pressure defibrillation contributed a unique
morphology of the interconnected web-like structure of nanofibers.
Chemical analysis of the cellulose fiber after each stage of purification showed an
increase in cellulose content and a decrease in lignin and hemicellulose content. Successive
bleaching helped with the cellulose purification. FTIR graph indicated the partial removal
of the pectins during the fiber extraction. It was also seen that the relative crystallinity of the
hemp fibers increased after each stage of chemical treatments.
Acknowledgment We gratefully acknowledge financial support of this study given by NSERC (Natural
Sciences and Engineering Research Council of Canada).

102

Appl Compos Mater (2007) 14:89103

References
1. Bhatnagar, A., Sain, M.: Processing of cellulose nanofiber-reinforced composites. J. Reinf. Plast.
Compos. 24, 12591268 (2005)
2. Chakraborty, A., Sain, M., Kortschot, M.: Reinforcing potential of wood pulp-derived microfibres in a
PVA matrix. Holzforschung 60(1), 5358 (2006)
3. Eichhorn, S.J., Baillie, C.A., Zafereiropoulos, N., Mwaikambo, L.Y., Ansell, M.P., Dufresne, A.,
Entwistle, K.M., Herrera-Franco, P.J., Escamilla, G.C., Groom, L., Hughes, M., Hill, C., Rials, T.G.,
Wild, P.M.: Review-current international research into cellulosic fibres and composites. J. Mater. Sci. 36,
21072131 (2001)
4. Chakraborty, A., Sain, M., Kortschot, M.: Cellulose microfibrils: a novel method of preparation using
high shear refining and cryocrushing. Holzforschung 59(1), 102107 (2005)
5. Nakagaito, A.N., Yano, H.: The effect of morphological changes from pulp fiber towards nano-scale
fibrillated cellulose on the mechanical properties of high-strength plant fiber based composites. Appl.
Phys., A 78, 547552 (2004)
6. Nakagaito, A.N., Yano, H.: Novel high-strength biocomposites based on microfibrillated cellulose
having nano-order-unit web-like network structure. Appl. Phys., A 80, 155159 (2005)
7. Sain, M., Bhatnagar, A.: Manufacturing of nanofibrils from natural fibres, agro based fibres and root
fibres. CA Patent Appl. 2,437,616 (2003)
8. Hepworth, D.G., Bruce, D.M.: The mechanical properties of a composite of a composite manufactured
from non-fibrous vegetable tissue and PVA. Compos., Part A Appl. Sci. Manuf. 31, 283285 (2000)
9. Gindl, W., Keckes, J.: Tensile properties of cellulose acetate butyrate composites reinforced with
bacterial cellulose. Compos. Sci. Technol. 64, 24072413 (2004)
10. Dinand, E., Chanzy, H., Vignon, M.R.: Suspensions of cellulose microfibrils from sugar beet pulp. Food
Hydrocoll. 13, 275283 (1999)
11. Canadian hemp alliances (2006) http://www.hemptrade.ca/en/public/about-hemp.ihtml. Cited 13 Aug
2006
12. Saheb, D.N., Jog, J.P.: Natural fiber polymer composites: a review. Adv. Polym. Technol. 18, 351363
(1999)
13. Oksman, K., Mathew, A.P., Bondeson, D., Kvien, I.: Manufacturing process of cellulose whiskers/
polylactic acid nanocomposites. Compos. Sci. Technol. 66(15), 27762784 (2006)
14. McCann, M.C., Wells, B., Roberts, K.: Direct visualization of cross-links in the primary plant cell wall.
J. Cell Sci. 96(2), 323334 (1990)
15. McCann, M.C., Wells, B., Roberts, K.: Complexity in spatial localization and length distribution of plant
cell wall matrix polysaccharides. J. Cell Sci. 166, 123136 (1993)
16. Clowes, F.A.L., Juniper, B.E.: Plant cell. In: Burnett, J.H., Phil, M.A.D. (eds.) vol. 8, pp. 207209.
Blackwell Scientific Publications, Oxford (1968)
17. Stamboulis, A., Baillie, C.A., Peijs, T.: Effects of environmental conditions on mechanical and physical
properties of flax fibres. Compos., Part A Appl. Sci. Manuf. 32, 11051115 (2001)
18. Dinand, E., Chanzyi, H., Vignon, M.R., Maureaux, A., Vincent, I.: Microfibrillated cellulose and method
for preparing a microfibrillated cellulose. US Patent 5,964,983 (1999)
19. Yoshinaga, F., Tonouchi, N., Watanabe, K.: Research progress in production of bacterial cellulose by
aeration and agitation culture and its application as a new industrial material. Biosci. Biotechnol.
Biochem. 61, 219224 (1997)
20. Krieger, J.: Bacterial cellulose near commercialization. Chem. Eng. News 68, 3537 (1990)
21. Bondeson, D., Kvien, I., Oksman, K.: Optimization of the Isolation of nanocrystals from microcrystalline
cellulose by acid hydrolysis. Cellulose 13, 171180 (2006)
22. Wan, W.K., Hutter, J.L., Millon, L., Guhados, G.: Bacterial cellulose and its nanocomposites for
biomedical applications. In: Oksman, K., Sain, M. (eds.) Cellulose Nanocomposites, pp. 221241.
Oxford University Press, Washington, DC (2006)
23. Ishihara, M., Yamanaka, S.: Modified bacterial cellulose. US Patent 6,627,419 (2002)
24. Oksman, K., Mathew, A.P., Bondeson, D., Kvien, I.: Manufacturing process of cellulose whiskers/
polylactic acid nanocomposites. Compos. Sci. Technol. 66(15), 27762784 (2006)
25. Taniguchi, T., Okamura, K.: New films produced from mircofibrillated natural fibres. Polym. Int. 47,
291294 (1998)
26. Mustat, A.: Factors influencing fiberfiber friction in the case of bleached flax. Cellul. Chem. Technol.
31, 405413 (1997)
27. Kvien, I., Tanem, B.S., Oksman, K.: Characterization of cellulose whiskers and their nanocomposites by
atomic force and electron microscopy. Biomacromolecules 6, 31603165 (2005)

Appl Compos Mater (2007) 14:89103

103

28. Kirby, A.R., Gunning, A.P., Waldron, K.W., Morris, V.J., Ng, A.: Visualization of plant cell walls by
atomic force microscopy. Biophys. J. 70, 11381143 (1996)
29. van der Wel, N.N., Putman, C.A.J., van Noort, S.J.T., de Grooth, B.G., Emons, A.M.C.: Atomic force
microscopy of pollen grains, cellulose microfibrils, and protoplasts. Protoplasma 194, 2939 (1996)
30. Thimm, J.C., Burritt, C.J., Ducker, W.A., Melton, L.D.: Celery (Apium graveolens L.) parenchyma cell
walls examined by atomic force microscopy: effect of dehydration on cellulose microfibrils. Planta 212,
2532 (2000)
31. Shimbun, S.K.: Senshoku nouhau no rironka [A Theorisation of Dyeing Know-How]. Japan (1985)
32. Annergren, G., Boman, M., Sandstrm, P.: Principal of multi-stage bleaching of softwood kraft pulp. In:
Proceedings of 1998 international pulp bleaching conference, Helsinki, Finland, 15 June (1998)
33. Sain, M., Bhatnagar, A.: Manufacturing of nano-sized fibers from renewable feedstock. US Patent Appl.
60,512,912 (2004)
34. Sain, M.: Solid phase dispersion and processing of micro- and nano-cellulosic fibres in plastic phase to
manufacture bio-nanocomposites products of commercial interests. CA Patent Appl. 2,559,844 (2006)
35. Zobel, B.J., Stonecypher, R., Browne, C., Kellison, R.C.: Variation and inheritance of cellulose in
southern pines. Tappi 49, 383387 (1966)
36. Stamm, A.J.: Wood and Cellulose Science. Ronald, New York (1964)

You might also like