You are on page 1of 21

This article was downloaded by: [University of Winnipeg]

On: 02 September 2014, At: 02:53


Publisher: Taylor & Francis
Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered
office: Mortimer House, 37-41 Mortimer Street, London W1T 3JH, UK

Composite Interfaces
Publication details, including instructions for authors and
subscription information:
http://www.tandfonline.com/loi/tcoi20

Characterization of new cellulose


sansevieria ehrenbergii fibers for
polymer composites
a a b
T.P. Sathishkumar , P. Navaneethakrishnan , S. Shankar & R.
a
Rajasekar
a
Department of Mechanical Engineering , Kongu Engineering
College , Erode , Tamilnadu , India
b
Department of Mechatronics Engineering , Kongu Engineering
College , Erode , Tamilnadu , India
Published online: 17 Jul 2013.

To cite this article: T.P. Sathishkumar , P. Navaneethakrishnan , S. Shankar & R. Rajasekar (2013)
Characterization of new cellulose sansevieria ehrenbergii fibers for polymer composites, Composite
Interfaces, 20:8, 575-593, DOI: 10.1080/15685543.2013.816652

To link to this article: http://dx.doi.org/10.1080/15685543.2013.816652

PLEASE SCROLL DOWN FOR ARTICLE

Taylor & Francis makes every effort to ensure the accuracy of all the information (the
“Content”) contained in the publications on our platform. However, Taylor & Francis,
our agents, and our licensors make no representations or warranties whatsoever as to
the accuracy, completeness, or suitability for any purpose of the Content. Any opinions
and views expressed in this publication are the opinions and views of the authors,
and are not the views of or endorsed by Taylor & Francis. The accuracy of the Content
should not be relied upon and should be independently verified with primary sources
of information. Taylor and Francis shall not be liable for any losses, actions, claims,
proceedings, demands, costs, expenses, damages, and other liabilities whatsoever or
howsoever caused arising directly or indirectly in connection with, in relation to or arising
out of the use of the Content.

This article may be used for research, teaching, and private study purposes. Any
substantial or systematic reproduction, redistribution, reselling, loan, sub-licensing,
systematic supply, or distribution in any form to anyone is expressly forbidden. Terms &
Conditions of access and use can be found at http://www.tandfonline.com/page/terms-
and-conditions
Downloaded by [University of Winnipeg] at 02:53 02 September 2014
Composite Interfaces, 2013
Vol. 20, No. 8, 575–593, http://dx.doi.org/10.1080/15685543.2013.816652

Characterization of new cellulose sansevieria ehrenbergii fibers for


polymer composites
T.P. Sathishkumara*, P. Navaneethakrishnana, S. Shankarb and R. Rajasekara
a
Department of Mechanical Engineering, Kongu Engineering College, Erode, Tamilnadu, India;
b
Department of Mechatronics Engineering, Kongu Engineering College, Erode, Tamilnadu, India
(Received 26 April 2013; accepted 15 June 2013)
Downloaded by [University of Winnipeg] at 02:53 02 September 2014

Natural cellulose fibers were newly identified from the sources of sansevieria ehren-
bergii plant. These fibers were extracted using the mechanical decortication process.
The hierarchical cell structure of the plant and fibers was analyzed using scanning
electron microscope, optical microscope, Fourier transforms infrared, and X-ray dif-
fraction. The density and diameter of the fibers were found to be approximately
0.887 g/cm3 and 10–250 μm, respectively. The various chemical compositions were
analyzed and compared with other natural fibers. The thermal stability of the fiber
was examined through thermogravimetric analysis/differential thermogravimetric
analysis (DTG). The maximum peak temperature was obtained at 333.02 °C in DTG
curve. The raw fibers exhibited a tensile strength of 50–585 MPa, an elongation at
break of 2.8–21.7%, a Young’s modulus of 2.5–7.5 GPa, and a corrected
compliances Young’s modulus of 2.5–7.8 GPa.
Keywords: sansevieria ehrenbergii; chemical composition; infrared spectroscopy;
X-ray diffraction; thermal stability

1. Introduction
Over the past two decades, polymer-based materials reinforced with various natural
fibers have been established everywhere from academic and industrial point of view.
Environmental consciousness, legislation, and energy consumption have inspired aca-
demics and industrial researchers working in the area of cellulose fibers and fiber-rein-
forced composite. More than 1000 species of cellulose plants are being available in
fibers forms and few of them are investigated to prepare the reinforced composite. The
natural fiber composites have attractive features likes low cost, light-in weight, moder-
ated strength, high specific modulus, moderate mechanical properties, easy to handle,
and lack of health hazards compared to synthetics fiber composite. Structure of the fiber
is framed with natural chemicals such as cellulose, hemicellulose, lignin, pectin, wax,
pentosen, and silica. The cellulose fiber-reinforced composites have been significantly
used for industrial components, construction material, automobile parts, and home appli-
ances. The traditionally available natural fibers have been used for making ropes, bed
sheet, covers, clothes, etc. These cellulose fibers, such as petiole bark, rachis, rachilla,
spatha, root, palmyrah, talipot,[1,2] sisal,[3–5] sansevieria cylindrica,[6] sea grass,[7]
coconut tree leaf sheath,[8] bamboo,[9] coir and banana,[5,9] curaua,[9,10] vakka, date,

*Corresponding author. Email: tpsathish@kongu.ac.in

Ó 2013 Taylor & Francis


576 T.P. Sathishkumar et al.

palm,[11] okra,[12] elephant grass,[13] pineapple leaf fiber,[2,14] abaca leaf fiber, san-
sevieria rifasciata, sisal leaf, cocount husk, kenaf bast, ramie bast,[14] rice husk and
phormium tenax,[15] kenaf,[15,16] wheat straw,[17] cotton,[18] flax and hemp,[15,18]
jute,[18,19] abaca, bagasse, alfa, piassava, date palm, henequen, isora, oil palm [9,19]
and softwood kraft [20] were extracted using various fiber-retting process, and also the
structural morphology of natural plants and fibers was analyzed using scanning electron
microscopy (SEM). The chemical compositions, and physical, thermal, and mechanical
properties of natural fibers were discussed with Fourier transforms infrared (FTIR), X-
ray diffraction (XRD), thermogravimetric analysis (TGA)/differential thermogravimetric
(DTG), and tensile testing. These fibers have been reinforced with polymer matrix to
prepare the composites using various manufacturing process.
In the present study, the newly identified sansevieria ehrenbergii fibers (SEFs) are
extricated from sansevieria ehrenbergii (SE) plant. The physical, chemical, mechanical,
Downloaded by [University of Winnipeg] at 02:53 02 September 2014

and thermal properties have been investigated through FTIR, XRD, TGA/DTG, and ten-
sile testing. The effects of fiber lengths on tensile properties such as strength, modulus,
and elongation at break, corrected compliances Young’s modulus, and Weibull modulus
have been investigated.

2. Natural fiber materials


Natural plant of SE commercially called as snake grass plant belongs to the family of
dracaenaceae. It is a flowering plant which grows in Northeastern Africa, Tanzania,
Saudi Africa, and South India and it was discovered by Schweinf Ex Baker in Olduvai
Gorge. It has been traditionally used for various purposes such as leaf for antiseptic and
natural bandage, fibers for rope-making, baskets, roofs, and clothes, and it is used as a
home garden at Vijayamangalam (Figure 1).

2.1. Structure of SE leaves and its fiber


SE is known as Blue sansevieria, Sword sansevieria, Oldupai or East African Wild
Sisal. It is a catus and succulent plant, and a native of Olduvai Gorge in Northern Tan-
zania. The structure of this plant consists of four to six leaves in dark green color and

Figure 1. SE natural plant.


Composite Interfaces 577

approximately 100–200 fibrils in a bundle. It grows in a fan shape with long stiff leaves
wrapped nearly into a sword pattern with sharp spine ending. These leaves are about
500–1500 mm long and 20–40 mm wide, and they grow in tight clusters in closer prox-
imity amongst themselves. A leaf consists of a sandwich natural composite containing
approximately 8% fiber, 1% cuticle, 12% dry matter, and 81% water.

2.2. Extraction process for SEFs


The natural plant fibers are extracted by adopting various retting process like mechani-
cal retting (by hammering or decorticator), chemical retting (boiling with chemicals),
steam/vapor retting, and water/microbial retting process.[4,8,13] The SEFs are extracted
from the SE leaves and they yield strong white elastic long fibers which have been used
for many purposes. In the process of fiber extraction, leaves of SE are plucked manu-
Downloaded by [University of Winnipeg] at 02:53 02 September 2014

ally and the nodes and culms of the leaves are trimmed by knife. The sand particles are
removed by water.
The mechanical decorticator consists of two rollers. The gap between these rollers
of 5 mm has been maintained for fiber extraction. The leaves are fed in between the
rotating rollers. The fibers are found to be adhering to the SE stem. The external layers
of fibers such as gums and skin of stems are removed by continuous feeding of leaves
in between the rotating rollers. The fibers separated from the stem are washed thor-
oughly with water and dried under the sun light for 48 h [21] to remove water content
from the fibers surface. These fibers are denoted as SEFs. The extracted fibers length is
around 500–1500 mm and it can be chopped into various lengths for experimentation.
This extraction is a very simple process; it requires less manpower and energy con-
sumption; and is very economical.

2.3. Physical analysis of SE plant and its fiber


The aim of this section is to analyze the cross-section of SE leaves and fibers, and their
microstructural characteristics and density of the fiber. SEM and optical microscope
(OP) are used to analyze the cross-section of the SE leaves and fibers. The image of
primary and secondary wall of the leaf and fibers is taken for calculating the dimension.
The cross-sectional area of the fibers is approximately round in shape and is used to
calculate the strength of the fiber. The average cross-section diameter is calculated from
the 1000 mm fiber length which is cut into 10 sections of 100 mm each.
The density of the SEFs is obtained using pycnometer for solids with toluene
(known as Toulon) as the immersion liquid.[13] Initially, the fibers are dried for 48 h in
air tight nonhygroscopic desiccator containing calcium chloride. Before measuring the
density of the fibers, they are initially impregnated in toluene for 2 h to remove the
microbubbles present in the fibers. The density of the toluene (qT Þ is 0.8669 g/cm3 at
20 °C. The fibers are chopped into length of 10 mm and kept into the pycnometer. The
following formula is used to calculate the SEF density (qSEFs Þ.[11]
 
m2  m1
qSEFs ¼ q ð1Þ
ðm3  m1 Þðm4  m2 Þ T

where m1 is the mass of the empty pycnometer (kg), m2 is the mass of the pycnometer
filled with chopped fibers (kg), m3 is the mass of the pycnometer filled with toluene (kg),
and m4 is the mass of the pycnometer filled with chopped fibers and toluene solution (kg).
578 T.P. Sathishkumar et al.

2.4. Chemical analysis of the SEFs


2.4.1. Cellulose content
The cellulose content of plant fiber has been measured by Kurshner and Hoffer’s
method.[6,22] The bundle of fibers was washed in distilled water for removing the
impurities present on the fibers surface and dried in an oven at 80 °C for 24 h. The sam-
ples of crushed fibers (150 mg) were hydrolyzed with a mixture of ethanol and 95%
nitric acid. This bundle of fibers was taken from the mixture and kept in oven for dry-
ing at 60 °C until constant weight. The cellulose content with a corresponding insoluble
fraction of samples was weighted. The hemicellulose was measured according to stan-
dard NFT 12-008.

2.4.2. Lignin content


Downloaded by [University of Winnipeg] at 02:53 02 September 2014

The insoluble solution of the lignin content of the SEFs was determined based on
Klason lignin of APPITA P11s-78 method.[22,23] The samples of crushed fibers
(300 mg) were hydrolyzed with 72% of sulfuric acid in an ultrasonic bath for 1 h at a
controlled temperature of 30 °C. The treated sample was mixed with a solvent solution
of methylene chloride (CH2Cl2) and placed in autoclave at 125 °C for one hour and
dried. Then, sample was cooled and the lignin was filtered through retentate. The
percentage of lignin content was obtained by following expression:

WLignin
IL ¼  100 ð2Þ
WFiber

where IL is the insoluble lignin content (%), WLignin is the oven dry weight of the insol-
uble lignin or Klason lignin (g), and WFiber is the oven weight of the fiber (g).

2.4.3. Wax content


The SEFs wax content was determined by Conrad method with Soxhlet extraction.[6]
The received SEFs were preconditioned before wax extraction took place. The bundle
of fibers was washed in distilled water for removing the impurities in the fibers surface
and dried in an oven at 80 °C for 24 h. Then, the chopped fiber of 10 mm was kept in
ethanol solution for 6 h. The solution-wetted fiber was transferred to separator funnel.
After that, the chloroform was added for wax extraction from the substance such as
waxes, fats, resins, photosterols and nonvolatile hydrocarbons, low-molecular-weight
carbohydrates, salts, and other water-soluble substance. The purified water was added
for separating the chloroform and other substances. Then, the chloroform was evapo-
rated by heating the funnel and leaving the wax. The percentage of wax content was
measured by following expression:

 
WFiber  WWax
%WAX ¼  100 ð3Þ
WFiber

where %WAX is the percentage of wax content, WFiber is the oven weight of the fiber
(g), and WWax is the weight of the wax extracted from the raw fiber.
Composite Interfaces 579

2.4.4. Ash content


Initial weight of the sample fibers was measured for quantifying the ash content. The
fibers were kept inside the colorimeter at 250 °C. The result of heating was obtained as
fiber ash. The percentage of ash content was measured by weighting the sample.

2.4.5. Percentage of moisture content


The moisture content of the fiber was obtained by normal weight loss method. Twenty
grams of fibers were weighted and kept in the oven to dry at 100 °C. During drying
process, the weight of the fibers bundle was measured until the constant weight was
obtained. The constant weight was obtained at 100 °C for 4 h 30 min. The following
expression was used to calculate the percentage of moisture content in the fibers.[22]
Downloaded by [University of Winnipeg] at 02:53 02 September 2014

 
WBefore drying  WAfter drying
%M ¼  100 ð4Þ
WBefore drying

where %M is the percentage of moisture content, WBefore drying is the weight of the fiber
before drying (g), and WAfter drying is the weight of the fiber after drying (g).

2.5. FTIR test


FTIR spectroscopy analysis of the SEFs was recorded using a Perkin Elmer Spectrum
RXI FTIR spectrometer. The spectrum was identified with scanning rate of 30 scans per
minutes. The total wave number region was recorded from 500 to 4000 cm1 at room
temperature of 30 °C and relative humidity of 65%. The bundle of fibers was chopped
approximately 2–3 mm and ground into powder form using mortar and pounder. The
powder was mixed with KBr matrix and pelletized using pressurization. The pellet was
used to record the spectra.

2.6. XRD analysis


The wide angle XRD spectra of the SEFs was recorded with a Rigaku X-ray diffrac-
tometer Dmax 2500. An X-ray tube produced monochromatic Cu Kα radiation. This
system had a rotating anode generator with angle powder goniometer. It was operated
at 40 kV and 150 mA. The reflection mode at a scan speed of 4 °C/min with high inten-
sity spectrum of SEF was identified and it estimated the crystallinity indices. All sam-
ples were scanned in 2h angle of 5–50°.

2.7. Thermogravimetric analysis


The thermograms of the steam SEFs were recorded on a Perkin Elmer TGA-7 instru-
ment in nitrogen atmosphere at a heating rate of 10 °C/min in the temperature range of
50–200 °C. About 10 mg of the sample fiber was heated from 30 to 200 °C at a rate of
5 °C/min, followed by an isothermal setup to 200 °C for 20 min. The flow rate of nitro-
gen gas was 30 mL/min. During this test, the weight loss in terms of temperature was
recorded as graphs. This test was obtained at controlled atmospheric condition with
65% RH and 30 °C.
580 T.P. Sathishkumar et al.

2.8. Single-fiber tensile test


The SEFs were chopped into various lengths for testing the tensile properties. Accord-
ing to the ASTM D3822-01,[6] the tensile test of single fibers was performed by uni-
versal tensile test machine of INSTRON 5500R in South Indian Textile Research
Association (SITRA), Coimbatore, Tamil Nadu, India. Twenty samples of single fiber
were tested for each fiber gage length such as 10, 20, 30, 40, and 50 mm. In testing,
the individual fiber was clamped with a pneumatic gripper, the cross-head speed of fiber
testing was 0.1 mm/min and the strain rates were on the order of 0.6–0.15/s. A 1.0 kN
load cell was used to measure the load vs. displacement curve (tensile strain) of fiber at
various gage lengths and dates were recorded as graph. The average value of the tensile
strength, strain, modulus, and elongation at break was obtained. All tests were con-
ducted at ambient temperature of 21 °C and humidity of about 65%. Finally, the
recorded dates were taken for analysis to calculate the tensile properties of fiber.
Downloaded by [University of Winnipeg] at 02:53 02 September 2014

The testing compliance of the loading and gripping systems is calculated by obtain-
ing the force vs. displacement behavior of the fiber at various gage lengths.[4,24] The
total cross-head displacement (δt) during fiber testing is expressed using the following
expression:
 
dt 1
¼ LþC ð5Þ
F EA

where F is the applied force, L is the gage length (GL), E is the Young’s modulus of
the fiber, A is the cross-sectional area of the fiber, and c is the machine compliance. A
plot of δt/F vs. GL yields a straight line of slope 1/(EA) and intercept in y axis, which
is the compliance (c) of the load train.

2.9. SEM and OP


The fibers cross-section in longitudinal and transverse directions was obtained using
SEM and OP. The image of transfer section of SE leaf was taken by a SEM model of
JEOL JSM-6390. The following specifications were used for scanning the image: (a)
resolution (3.0 nm (Acc V 30 kV, WD 8 mm, and SEI)), (b) magnification (5X (WD
48 mm or less)), and (c) electron gun (accelerating voltage: 0.5–30 kV and filament:
pre-centered tungsten hairpin filament). The cross-area and diameters were obtained by
these images.

3. Results and discussion


3.1. Macrostructural analyses of SEF leaf and its fibers
Figure 2 shows that the transverse section of SE leaf consists soft and hard tissues, and
the fibers are arranged in the soft tissue of the plant (Figure 2(A) and (B)). Fibers are
grown from base of the plant which is classified as structural and arch fiber. The struc-
tural fibers provide strength and stiffness for the SE leaf and are found in the periphery
of the leaf (Figure 2(A)). The cross-section of the structural fibers is found in semi-
elongated circular shape. The arch fibers are usually found in the middle of the leaf
(Figure 2(B)). The structure fibers present in the middle of the plant are of higher diam-
eter and also found more in numbers. The structure fibers are nearer to green layer of
the plant (Figure 2(C)). The diameter of the fibers is small and they are found less in
Composite Interfaces 581
Downloaded by [University of Winnipeg] at 02:53 02 September 2014

Figure 2. SEM image of cross-section of SE leaf and fiber: (A) and (B) middle of the SE plant
at transfer section; (C) and (D) image of near to green layer; and (E) and (F) the longitudinal
cross-section of fiber.

numbers. The structural fibers are very important because they never split during
the extraction process. The skin (green layer) of the plant does not have fiber
(Figure 2(D)).

3.2. Physical properties of fiber


Some literature reports assumed the fiber are in circular cross-section. The SEFs have a
semi-elongated circular shape, the OP is used to measure the diameter of the fiber (10
samples) at four points with equal spacing through a 100 mm of 1000 mm-long fibers.
The fiber diameter is almost 20–250 μm. But, the image analysis from an SEM
582 T.P. Sathishkumar et al.
Downloaded by [University of Winnipeg] at 02:53 02 September 2014

Figure 3. The optical image of the longitudinal section of SEFs.

micrograph is used to measure the area of fiber in precise method. The area of the fiber
has measure at four points with equal spacing through a 40 mm SEFs which resulted in
values of 0.012, 0.016, 0.018, and 0.0215 mm2 respectively. Therefore, measuring the
area of tensile-tested fiber at the fractured end resulted in more accurate determination
of the average area. For example, the area of the fractured end was 0.011, 0.0141,
0.017, and 0.020 mm2 respectively.
Figure 3 shows the longitudinal section of SEFs. The cross-sections of fibers
are measured by OP along the length of the fiber at various locations. The fiber
cross-sections varied with varying fiber length. The images of the fiber are
postprocessor to Image ‘J’ software for determining the fiber diameters (20–240 μm).
The density of various natural fibers is given in Table 1. The density may vary with
various fibers which are due to various climatic conditions, plant growth rate, and plant
tissue. The density of SEF is 0.887 g/cm3 as measured. The SEF has less density com-
pared to sisal, flax, oil palm, hemp, jute, sea grass, agave fiber, etc. and more density
than petiole bark, elephant grass, abaca leaf, elephant grass, etc. The weight of the com-
posites can be determined by the fiber density. The low density fiber has been rein-
forced with matrix, which leads to less weight composite. The SEF is a low density
fiber. SEF-reinforced polymer composites can be used for light-weight applications.
This low density is in the construction of thermal and acoustic insulation using panels
reinforced with SEFs.

3.3. Analysis of chemical composition of fiber


The chemical composition of the fiber influences its properties. The cellulose content
has an important influence on the mechanical properties of the natural fiber such as ten-
sile strength, modulus, and elongation at break. The various chemical compositions of
various natural fibers are listed in Table 2. The SEFs contain 80% cellulose, which is
2.5, 7.5, 12, 12.7, 18.8, 29, and 53.8% greater than sisal, hemp, jute, flax, oil palm,
kenaf and coir fibers, which is 3.8, 11.11% less than banana and cotton which is equal
to sansevieria cylindrica and pineapple leaf fibers.
The SEFs also contain 11.25% of hemicellulose, 7.8% of lignin, 0.45% of wax, and
10.55% of moisture. The hemicellulose of SEFs is less compared to all other fiber. The
lignin content influences the fiber structure, properties, and morphology. It is less than
Composite Interfaces 583

Table 1. Physical and tensile properties of raw SEF and various natural fiber.

Gage Tensile Tensile


Density Diameter length strength modulus Elongation
Fiber name (g/cm3) (μm) (mm) (MPa) (GPa) (%)
Petiole bark [1,2] 0.69 250–650 – 185.52 15.09 2.1
Spatha [1,2] 0.69 150–400 – 75.66 3.14 6
Vakka [11] 0.81 175–230 100 549 15.85 3.46
Elephant grass [13] 0.817 70–400 50 185 7.40 2.5
Abaca leaf fiber 0.83 114–130 10 418–486 12–13.8 –
[14]
Sansevieria 0.887 20–250 10–50 50–585 1.5–7.67 2.8–21.7
ehrenbergii
Sansevieria 0.89 83–93 10 526–598 13.5–15.3 –
rifasciata [14]
Downloaded by [University of Winnipeg] at 02:53 02 September 2014

Bamboo [9] 0.91 240–330 100 503 0.35–0.91 1.40


Sansevieria 0.915 230–280 10–40 585–676 0.2–11.2 11–14
cylindrica [6]
Date [11] 0.99 155–250 100 309 11.32 2.73
Palmyrah [1,2] 1.09 20–80 43 180–215 7.7–20 7–15
Palm [11] 1.03 400–490 100 377 2.75 13.71
Coconut or coir [8] 1.15 100–450 100 500 2.5 20
Agave [24] 1.20 126–344 – – – –
Henequen [9,19] 1.20 – – 430–570 11.1–16.3 3.7–5.9
Bagasse [20] 1.25 200–400 – 290 11 –
kenaf bast [14] 1.31 65–71 10 427–519 23.1–27.1 –
Pineapple leaf 1.32 53–62 10 609–700 23.7–30.3 0.8–1.6
[2,14]
Banana [5,9] 1.35 60–250 150 600 17.85 3.36
Curaua [9,10] 1.40 170 10 158–729 – 5
Sisal [3,4] 1.45 50–300 50 530–640 9.4–22 3–7
Hemp [15,18] 1.48 – – 550–900 70 2–4
Jute [18] 1.3–1.5 40350 60 393–773 10–30 1.5–1.8
Flax [15,18] 1.4–1.5 – – 345–1500 27.6–80 0.2–3.2
Sea grass [7] 1.50 5 – 453–692 3.1–3.7 13–26.6
Oil palm [9,19] 1.55 – 248 3.2 25

kenaf, jute, oil palm, pineapple leaf fibers, and some other fibers, and also greater than
flax, sea grass, sansevieria cylindrica, banana, agave, and some other fibers. The wax
content in raw fiber affects the interfacial bond between the fiber and the matrix during
composition preparation. However, SEFs contain less wax content than kenaf, jute,
hemp, phormium tenax, nettle, and oil palm fibers, and it is greater than agave and san-
sevieria cylindrica fiber. Therefore, to improve the interfacial bonding between the SEFs
and resin, chemical treatment of the fiber has to be done in order to remove the wax
and moisture content.

3.4. Fourier transform infrared spectroscopy


Figure 4 shows the FTIR spectrum of wave number 1500–400 cm1 of raw SEF. The
FTIR spectrum of SEF shows absorption bands of various chemical groups of lignocellu-
lose fiber components such as cellulose, hemicellulose, and lignin. The various principal
components are alkenes, phenolic hydroxyl group, aromatic groups, β-glucosi linkages,
and various functional groups containing oxygen (ester, ketone and alcohol).[25]
584 T.P. Sathishkumar et al.

Table 2. Chemical compositions of raw SEF and various natural fibers.

Hemi
Cellulose Lignin cellulose Pectin Wax Moisture Ash
Fiber name (%) (%) (%) (%) (%) (%) (%)
Sansevieria 80 7.8 11.25 – 0.45 10.55 0.6
ehrenbergii
Kenaf [14] 45–57 21.5 8–13 0.6 0.8 6.2–12 2–5
Jute [18] 61–71.5 11.8–13 17.9–22.4 0.2 0.5 12.5–13.7 0.5–2
Hemp [15,18] 70.2–74.4 3.7–5.7 17.9–22.4 0.9 0.8 6.2–12 0.8
Phormium tenax [15] 45.1–72.0 11.2 30.1 0.7 0.7 10 –
Flax [15,18] 64.1–71.9 2.0–2.2 64.1–71.9 1.8–2.3 1.7 8–12 –
Agave [24] 68.42 4.85 4.85 – 0.26 7.69 –
Bamboo [9] 26–43 1–31 30 – – 9.16 –
Sea grass [7] 57 5 38 10 – – –
Downloaded by [University of Winnipeg] at 02:53 02 September 2014

Sisal [3,4] 78 8 10 – 2 11 1
Sansevieria cylindrica 79.7 3.8 10.13 – 0.09 3.08 –
[6]
Oil palm [9,19] 65 17.5 10.12 – 4 – –
Henequen [9,19] 60 8 28 – 0.5 – –
Coconut tree leaf 27 27.7 14 – – 4.7 –
sheath [8]
Piassava [9] 28.6 45 25.8 – – – –
Cotton [18] 82.7–90 – 3 – 0.6 7.85–8.5 –
Rice husk [15] 38–45 – 12–20 – – – 20
Coniferous [22] 40–45 26–34 – – – – –
coir 37 42 – – – 11.36
Banana [5,9] 83 5 – – – 10.71

Figure 4. FTIR spectra of raw SEF.


Composite Interfaces 585

A broad absorption band in between 3600 and 3100 cm1 (attributed to cellulose Iβ) cor-
responds to the characteristic of O–H stretching vibration and hydrogen bond of the
hydroxyl groups.[26] The absorption bands between 3786 and 3457 cm1 belongs to the
stretching vibrations of the phenolic hydroxyl group in the lignin components and hydro-
xyl group of the glucopyranose unit in the cellulose. The two sharp peaks near at 2945
and 2855 cm1 are the characteristic bands for the C–H stretching vibration from CH
and CH2 in cellulose and hemicellulose components. A small absorption peaks at
1636 cm1 belong to C=O stretching vibration of the alpha keto carboxylic acid in lignin
or ester group in hemicellulose. The CH2 symmetric bending band is absorbed at a peak
of 1482 cm1.[24] This represents the hemicellulose xylene, wax, and impurities. The
peaks at 1385 cm1 belong to C–O stretching vibration of the acetyl group in lignin and
hemicellulose. The bending vibration C–H and C–O group of aromatic ring of hemicell-
ulos and lignin shows peaks at 1122 and 1035 cm1 stretched vibration.[6] Also, the β-
Downloaded by [University of Winnipeg] at 02:53 02 September 2014

glucosi linkages present in cellulose and hemicellulose which is absorbed at 840 cm1
stretching vibration.

3.5. TGA and DTG


The TGA/DTG vs. temperature curves of the raw SEF thermal process are shown in
Figure 5. The temperatures used for the analysis range from 30 to 500 °C. The weight
loss may be associated with dehydration as well as degradation of lignin content in
fiber. This process occurs in the temperature range from 50 to 230 °C. The temperature
range from 233 to 352 °C leads to higher weight loss which decomposes the cellulose.
Almost all cellulose content of the SEF is decomposed at the temperature of 360 °C.
Normally, the moisture content of the fiber is associated with the thermal stability of

Figure 5. TGA/DTG curve of raw SEF.


586 T.P. Sathishkumar et al.

the fiber. The weight loss may be associated with loss of water in the natural fiber.
[6,16,24] In DTG curve, the initial peak observed at 60 °C for SEF corresponds to the
heat of evaporation of water from the fiber. The second peak of the SEF observed at
333 °C is due to the thermal depolymerization of hemicellulose and the cleavage of the
glycosidic linkage of cellulose. In DTG curve, the third peak of the SEF observed at
450 °C is due to the thermal depolymerization of wax and other substance in the fiber.
This peak is associated for calefactive rate. The temperature of SEF in onset degrada-
tion and peak degradation is lower compared to coconut leaf sheath,[8] Okra,[12]
Kenaf, [16] and Agave americana.[24] The moisture content of okra, coconut leaf
sheath, and agave americana fiber is low compared to SEF. In raw SEF, the molecules
are not closely packed, which leads to more penetration of water in the fibers. This ulti-
mately decreases the thermal stability of the fiber. The higher value of moisture content
in SEF needs to pre-dry and chemically treat them before making composites.
Downloaded by [University of Winnipeg] at 02:53 02 September 2014

3.6. XRD analysis


The X-ray spectrum for raw SEF is shown in Figure 6. This spectrum shows two peaks
which are mostly distinct for a raw fiber. These two diffraction peaks indicate that the
SEF is semi-crystalline. The two peaks found at 2h = 14.26° and 2h = 26.3°. This can be
attributed to cellulose I and IV, both of which exhibit a monoclinic structure [18=26].
The broadening of the peak at 2h = 14.26° is mostly of noncellulose materials like
hemicellulose and lignin in the fiber. The peak at 2h = 26.3° is the content of α-cellulose
in the fiber.[22] These peaks are attributed to the (1 1 0) and (2 0 0) crystalline planes.
[26] The crystallinity index (CI) is measure of the crystalline cellulose using the follow-
ing expression:

Figure 6. XRD analysis of SEF.


Composite Interfaces 587

H26:3  H14:26
CI ¼ ð6Þ
H26:3

where H26.3 is the height of the peak at 2h = 26.3°. H14.26 is the height of the peak at
2h = 26.3 and it denotes the intensity of the amorphous peak in the crystallographic
planes. This calculation yields a CI value of 52.27%. It is higher than the cordia dicho-
toma fabric [22] of 12.14% and lower than the sansevieria cylindrica of 60%.[6]

3.7. Tensile properties of SEFs


A total of 20 fibers are randomly chosen from a given bundle and tested for each GL
of 10, 20, 30, 40, and 50 mm. Figure 7 shows the tensile testing machine. The fiber
ends are hold between the pneumatic grippers and the load vs. displacement curve is
Downloaded by [University of Winnipeg] at 02:53 02 September 2014

measured. Figure 8 shows the tensile force vs. displacements curve measured for single
fiber with varying GL. The increase in GL of fiber leads to more fiber displacement.
The 10 mm GL has less displacement and carries higher tensile load. Increasing the GL
decreases the load carrying capability of fiber. This is due to increase in the number of
flaw in the fiber. The fiber length exceeding 40 mm does not significantly change in

Figure 7. Single-fiber tensile testing machine.


588 T.P. Sathishkumar et al.

displacement. This may be due to constant size of mean flaw in the fiber. The Young’s
modulus is calculated in the elastic portion of the stress–strain curve and then corrected
for compliance by measuring the force vs. displacement at various gage lengths using
Equation (5) (Figure 9). These results are shown in Table 3. The GL of fiber does not
influence the modulus of the fiber. The variability in the modulus for a given GL is due
to the changeability in the microstructure of SEFs by various climatic conditions and
also due to possible damage that occurred during the mechanical decorticator process.
Downloaded by [University of Winnipeg] at 02:53 02 September 2014

Figure 8. Tensile force vs. displacements of various GL of fiber.

Figure 9. Displacement/forced of fibers vs. gage length (showing a linear relationship with a
slope of 1/(EA)).
Composite Interfaces 589

A measured vs. corrected stress–strain response of the SEFs is used to quantify the
machine compliance. At a given stress level, the measured machine compliance has
contributed in larger displacements. The displacement/force vs. GL is used to calculate
the machine compliance at two different levels. The slope of the line is also measured
for calculating the modulus of the fiber at different GL. The modulus calculated from
the corrected compliance curve is similar to the modulus calculated from the displace-
ment/force vs. GL procedure. The tested fiber result of a nonlinear region following the
initial portion of the stress strain curve is observed. This may happen due to a collapse
of the weak primary and secondary cell walls, and delamination between the microfiber
cells.[4] During the tensile testing of fiber, the strain-to-failure rate of the fibers
decreased with increasing the gage length of SEFs. This may be related to the size of
the flaws, propagation of flaws in the entire volume of the fiber, and the number of
microcrack due to GL increments.
Downloaded by [University of Winnipeg] at 02:53 02 September 2014

The tensile strength of the fiber does not depend on function of the GL as shown in
Table 3. The Young’s modulus varies between 2.5 and 7.5 GPa and the mean tensile
strength varies between 50 and 585 MPa. The tested parameters influencing the results
are as follows: precision of the instrumentation, gage length, strain rate, type of grips,
and compliance of the machine. Increase in gage length affects the ductility of the fiber.
Strength and modulus will not be affected as the mean flaws size controls the strength.
During testing, the crack propagation is initiated as the largest flaws determines the duc-
tility. The gage length increases the several flaws existed and the linkage between the
flaws will be quicker during testing. It is found that the ductility is very low in the
highest gage length of 50 mm. The mechanical behaviors of the SEFs are also depen-
dent on source of plant, climate conditions, growth rate, age, types of processing, and
the fiber microstructure.
The SEFs exhibited variability in tensile strength which is quite characteristic of all
natural fibers. This changeability can be explained by distribution of defects within the
fiber and fiber surface. Slight differences in microstructure associated with the different
types of SEFs could also result in some variability. Weibull statistic tool is used to rank
the relative fiber strength vs. probability of failure of the fibers to obtain a measure of
the variability in fiber strength.[4] According to the Weibull analysis, the probability of
survival of a fiber at a stress (σ) is given by the following relation:

  m 
r
PðrÞ ¼ exp  ð7Þ
ro

where σ is the fiber strength for a given probability of survival, and m is the Weibull
modulus. σo is defined as the characteristic strength. The higher the value of m, the
lower the variability in strength. Ranking defined as the characteristic strength of the
fiber strengths is performed by using an estimator given by:

i  0:5
Estimator ðE1 Þ ) PðrÞ ¼ 1  ð8Þ
N

i
Estimator ðE2 Þ ) PðrÞ ¼ 1  ð9Þ
N þ1
Downloaded by [University of Winnipeg] at 02:53 02 September 2014

590

Table 3. Summary of single-fiber tensile properties and Weibull modulus of various estimators of SEF.

Tensile strength Young’s modulus Young’s modulus corrected Area of the


Gage length (mm) Diameter (μm) (MPa) (GPa) for compliance (GPa) Strain-to-failure (%) Weibull modulus fiber mm2
10 60–240 317.65 ± 266.78 2.27 ± 1.13 2.57 ± 1.05 15.75 ± 5.93 7.28 7.29 7.28 7.32 0.024 ± 0.003
20 55–225 343.67 ± 242.99 3.29 ± 1.79 3.58 ± 2.11 9.05 ± 2.61 7.29 7.31 7.30 7.33 0.028 ± 0.005
30 40–230 347.12 ± 230.12 4.07 ± 2.56 5.45 ± 1.19 7.53 ± 2.97 7.15 7.16 7.15 7.18 0.025 ± 0.003
40 30–235 356.28 ± 267.27 5.55 ± 1.91 5.95 ± 1.17 6.81 ± 4.42 7.14 7.14 7.14 7.17 0.023 ± 0.006
50 35–245 368.18 ± 198.58 6.37 ± 2.15 6.56 ± 1.73 5.25 ± 2.42 7.07 7.08 7.07 7.11 0.019 ± 0.008
T.P. Sathishkumar et al.
Composite Interfaces 591

i  0:3
Estimator ðE3 Þ ) PðrÞ ¼ 1  ð10Þ
N þ 0:4

i  3=8
Estimator ðE4 Þ ) PðrÞ ¼ 1  ð11Þ
N þ 0:25

where P(σ)i term is the probability of survival corresponding to the ith strength value
and N is the total number of fibers tested (N = 20). Substituting Equations (8–11) into
Equation (7) yields:
 
r
In In½E ¼ m In ð12Þ
ro
 
Downloaded by [University of Winnipeg] at 02:53 02 September 2014

Thus, a plot of In In½E vs. In rro yields a straight line with slope of m. The Weibull
modulus for the SEF is higher than sisal fiber. Decreasing the Weibull modulus increases
the GL of fiber. The Weibull modulus obtained from estimators may be varied with small
changes. Silva et al. [4] obtained a Weibull modulus of 4.6–3.0 for 10–40 mm fiber
lengths for sisal fiber. Figure 10 shows the influence of GL on the Weibull modulus m
with estimator E2. Increasing the GL of SEFs resulted in a decrease in Weibull modulus.
It can be seen from Figure 8 that the curves have different slope – they almost all inter-
cept at the same point (e.g. they have the same mean strength, but little variation).
According to Silva et al. [4], this behavior of different materials has the same mean
strength and different Weibull modulus can be obtained. The mean strength of the all
materials can be controlled by mean defect size in the fiber, but the number of defects in
incremental fiber length controls the Weibull modulus. Thus, the fibers with a lower GL
(10 mm) have less number of defects and the average defect size is same as increasing

Figure 10. Weibull distribution of the SEFs tensile strength for different gage lengths of E2.
592 T.P. Sathishkumar et al.

the fiber GL. The defect meant the flaw in the fiber and fiber surface. However, the fiber
strength is not affected by the fiber GL, but the ductility is affected by GL. This means
that once a crack is formed at the largest flaw, how quickly this linkage between flaws
occurs determines the ductility of fiber. Increasing GL of SEFs leads to a large number
of flaws, the linkage between the flaws will be quicker which leads lower ductility (as
absorbed in the highest SEF GL of 50 mm).

4. Conclusion
The structure and properties of the SEFs were studied. The SEM analysis showed the
morphology of the plant and fibers was arranged in the plant. The diameter of the fiber
along longitudinal section was around 25–250 μm using optical microscopy and
40–165 μm using SEM. Also, the diameter of the fibers of 20–240 μm was obtained from
Downloaded by [University of Winnipeg] at 02:53 02 September 2014

transverse section of the fibers using optical microscopy. The cross-sectional area of one
these fibers was 0.0215 mm2. The physical density of the SE fiber was 0.887 g/cm3 which
has been used to prepare light-weight composites materials. The XRD and FTIR analyses
showed that SEFs were semi-crystalline and Iβ cellulose was present in the fiber. The ten-
sile strength of the fibers is independent of the gage length. The Young’s modulus of SEFs
corrected for machine compliance was approximately 7.8 GPa. The strain-to-failure
decreased from 2.8 to 21.7% when the gage length was increased from 10 to 50 mm. The
Weibull modulus was decreased with respect to different estimator and decreased when
the gage length was increased.

References
[1] Satyanarayana KG, Sukumaran K, Mukherjee PS, Pavithran C, Pillai SGK. Natural fiber –
polymer composite. Cem. Concr. Compos. 1990;12:117–136.
[2] Madhi E, Hamouda ASM, Sen AC. Quasi-static crushing behaviour of hybrid and non-
hybrid natural fiber composite solid cones. Compos. Struct. 2004;66:647–663.
[3] Li X, Tabil LG, Paanigrahi S. Chemical treatments of natural fiber for use in natural fiber–
reinforced composites: a review. J. Polym. Environ. 2007;15:25–33.
[4] Silva FA, Chawla N, Toledo Filho RD. Tensile behavior of high performance natural (sisal)
fibers. Compos. Sci. Technol. 2008;68:3438–3443.
[5] Joseph K, Toledo filho RD, James B, Thomas S, De Carvalho LH. A review on sisal fiber
reinforced polymer composites. Revista brasileira de engenharia agrícola e ambiental
1999;3:367–379.
[6] Sreenivasan VS, Somasundaram S, Ravindran D, Manikandan V, Narayanasam R. Micro-
structural, physico-chemical and mechanical characterisation of Sansevieria cylindrica
fibers – an exploratory investigation. Mater. Des. 2011;32:453–461.
[7] Davies P, Morvan C, Sire O, Baley C. Structure and properties of fibers from sea-grass
(Zostera marina). J. Mater. Sci. 2007;42:4850–4857.
[8] Obi Reddy K, Sivamohan Reddy G, Uma Maheswari C, Varada Rajul A, Madhusudhana
Rao K. Structural characterization of coconut tree leaf sheath fiber reinforcement. J. For.
Res. 2010;21:53–58.
[9] John MJ, Anandjiwala RD. Recent developments in chemical modification and characteriza-
tion of natural fiber-reinforced composites. Polym. Compos. 2008;29:187–207.
[10] Monteiro SN, Aquino RCMP, Lopes FPD. Performance of curaua fibers in pullout tests.
J. Mater. Sci. 2008;43:489–493.
[11] Murali Mohan Rao K, Mohana Rao K. Extraction and tensile properties of natural fibers:
vakka, date and bamboo. Compos. Struct. 2007;77:288–295.
[12] De Rosa IM, Kenny JM, Maniruzzaman M, Maniruzzaman Md, Puglia MD, Santulli C,
Sarasini F. Effect of chemical treatments on the mechanical and thermal behaviour of okra
(Abelmoschus esculentus) fibers. Compos. Sci. Technol. 2010;71:246–254.
Composite Interfaces 593

[13] Murali Mohan Rao K, Ratna Prasad AV, Ranga Babu MNV, Mohan Rao K, Gupta AVSSKS.
Tensile properties of elephant grass fiber reinforced polyester composite. J. Mater. Sci.
2007;42:3266–3277.
[14] Munawar SS, Umemura K, Kawai S. Characterization of the morphological, physical, and
mechanical properties of seven nonwood plant fiber bundles. J. Wood. Sci.
2007;53:108–113.
[15] De Rosa IM, Santulli C, Sarasini F. Mechanical and thermal characterization of epoxy com-
posites reinforced with random and quasi-unidirectional untreated Phormium tenax leaf
fibers. Mater. Des. 2010;31:2397–2405.
[16] Yussuf AA, Massoumi I, Hassan A. Comparison of polylactic acid/kenaf and polylactic
acid/rise husk composites: the influence of the natural fibers on the mechanical, thermal and
biodegradability properties. J. Polym. Environ. 2011;18:422–429.
[17] Le Digabel F, Averous L. Effect of ligno cellulose content on the properties of lignocellu-
lose-based biocomposites. Carbohydr. Polym. 2006;23:537–545.
[18] Venkateshwaran N, Elayaperumal A. Banana fiber reinforced polymer composites – a
Downloaded by [University of Winnipeg] at 02:53 02 September 2014

review. J. Reinf. Plast. Compos. 2010;29:2387–2396.


[19] Joseph S, Joseph K, Thomas S. Green composites from natural rubber and oil palm fiber:
physical and mechanical properties. J. Polym. Mater. 2006;55:925–945.
[20] Holber J, Houston D. Natural-fiber-reinforced polymer composites in automotive applica-
tions. JOM. 2006;58:80–86.
[21] Athijayaman A, Thiruchitrambalam M, Natarajan U, Pazhanivel B. Effect of moisture
absorption on the mechanical properties of randomly oriented natural fibers/polyester hybrid
composite. Mater. Sci. Eng. 2009;517:344–353.
[22] Jayaramudu J, Maity A, Sadiku ER, Guduri BR, Varada Rajulu A, Ramana CH, Li R. Struc-
ture and properties of new natural cellulose fabrics from Cordia dichotoma. Carbohydr.
Polym. 2011;86:1623–1629.
[23] Raiskila S, Pulkkinen M, Laakso T, Fagerstedt K, Loija M, Mahlberg R, Paajanen L,
Ritschkoff AC, Saran P. FTIR spectroscopic prediction of Klason and acid soluble lignin
variation in Norway spruce cutting clones. Silva. Fenn. 2007;41:351–371.
[24] Thamae T, Baillie C. Influence of fiber extraction method, alkali and silane treatment on the
interface of Agave Americana waste HDPE composites as possible roof ceilings in Lesotho.
Compos. Inter. 2007;14:821–836.
[25] Burgueno R, Quagliata MJ, Mehta GM, Mohanty AK, Misra M, Drzal LT. Sustainable cellu-
lar biocomposites from natural fibers and unsaturated polyester resin for housing panel appli-
cations. J. Polym. Environ. 2005;31:139–149.
[26] Saravana kumar SS, Kumaravel A, Nagarajan T, Sudhakar P, Baskaran R. Characterization
of a novel natural cellulosic fiber from Prosopis juliflora bark. Carbohydr. Polym.
2013;92:1928–1933.

You might also like