You are on page 1of 8

Composites: Part A 30 (1999) 277284

Some mechanical properties of untreated jute fabric-reinforced polyester


composites
T. Munikenche Gowda a, A.C.B. Naidu a,*, Rajput Chhaya b
a

B.M.S. College of Engineering, Bangalore-560019, India


Aeronautical Development Establishment, Bangalore, India
Received 10 December 1997; accepted 1 July 1998

Abstract
This research work is concerned with the evaluation of the mechanical propertiesmodulus, Poissons ratio and strengthof woven jute
fabric-reinforced composites. The specimens are prepared using hand lay-up techniques as per the ASTM standard. This is the first report by
any single group of researchers in which tensile strength, compressive strength, flexural strength, impact strength, inplane shear strength,
interlaminar shear strength and hardness are given. This work being an experimental study on untreated (as received jute fabric) woven jute
fabric-reinforced polyester composites, demonstrates the potential of this renewable source of natural fibre for use in a number of consumable
goods. q 1999 Published by Elsevier Science Ltd. All rights reserved.
Keywords: B. Mechanical properties; B. Strength; A. fibres; woven fabric

1. Introduction
Glass, carbon, boron and Kevlar fibres are being used as
reinforcing materials in fibre reinforced plastics (FRP)
which have been widely accepted as materials for structural
and non-structural applications. The main reason for the
interest in FRP is due to their high specific modulus, high
stiffness to weight ratio and high strength to weight ratio
compared to conventional materials. However, these materials are prohibitively expensive and their use is justified
only in aerospace applications. Therefore, natural fibres
like banana, cotton, coir, sisal and jute have attracted the
attention of scientists and technologists for application in
consumer goods, low-cost housing and other civil structures. It has been found that these natural fibre composites
possess better electrical resistance, good thermal and acoustic insulating properties and higher resistance to fracture.
Among all the natural fibre reinforcing materials jute
appears to be a promising material because it is relatively
inexpensive and commercially available in the required
form. It has higher strength and modulus than plastic [1]
and is a good substitute for conventional fibres in many
situations. However, the jute fibre has a multicellular structure composed of microfibrils and the cross-section is highly
non-uniform. It has been found that the mechanical and

physical properties are highly inconsistent and depend on


geographic origin, climatic growth conditions and processing techniques.
Bhattacharyya et al. [2] have studied the effect of process
variables such as curing temperature and time on the
mechanical properties of jute fibres in phenolformaldehyde.
Scientists at the Vikram Sarabhai Space Centre (Trivendrum) and the National Aerospace Laboratory (Bangalore)
have carried out considerable work on jute fibre in epoxy
polyester resins. However, their emphasis is more on
product development. Verma et al. [3] and Mohan et al.
[4] have studied the mechanical properties of jute/glass
hybrid composites in polyester resin and epoxy resin.
Chawla and Bastos [5] studied the effect of the volume
fraction of untreated jute fibres in unsaturated polyester
resin made by the leaky mould technique on Youngs modulus, maximum strength and impact strength. Winfield [6, 7]
studied the application of jute-reinforced plastics to low cost
housing. In this detailed study jute fabric is directly used as
a reinforcing material. Literature in hand reveals that no
single group of researchers has completely characterised
the mechanical properties of jute fabric-reinforced polyester
composites for (i) tensile strength and modulus, (ii)
compressive strength and modulus, (iii) flexural strength
and modulus, (iv) impact strength, (v) inplane shear strength
and modulus, (vi) interlaminar shear strength, and (vii)

* Corresponding author.
1359-835X/99/$ see front matter q 1999 Published by Elsevier Science Ltd. All rights reserved.
PII: S1359-835 X( 98)00 157-2

278

T. Munikenche Gowda et al. / Composites: Part A 30 (1999) 277284

hardness of untreated jute fabric-reinforced polyester


composites.
2. Processing and chemical composition of jute fibres
2.1. Jute fibre
Jute is a bast fibre obtained from inner bast tissues of the
plant stem [8]. The fibres are bound together by gummy
materials (pectinous substances) which keep the fibre
bundles cemented with non-fibrous tissues of jute bark.
These encircling soft tissues must be softened, dissolved
and washed away so that the fibre can be obtained from
the stem. This is done by steeping the stems in water and
is known as retting. The optimum water temperature for
retting is 808F. Micro-organisms (mainly bacillus bacteria)
decompose the gums and soften the tissues in 5 to 30 days
depending upon temperatures and the type of water used. It
has been found that the presence of higher amounts of
calcium and magnesium tend to increase the tenacity of
fibre.
As-received jute fibres will have very irregular transverse sections and a microcellular structure composed of
microfibrils. The fibre cross-section is highly irregular. In
addition, approximately one thousand varieties of jute have
been identified by the raw jute division of the Indian Jute
Institute Research Association (IJIRA), Calcutta. The biological division of IJIRA has been engaged in intensive study
of the improvement of retting. Thus, one has very little
control over the properties and characteristics of jute fibres.
The retting process also influences the strength, colour and
lustre of the fibre. There are two main types of retting (i)
manual and (ii) mechanical. During manual retting, the
fibres are freed from the stalk by tapping lightly with a
mallet and then lashing the stems in water. In mechanical
retting, a mechanical decorticator frees the fibres from the
stalk, greatly reducing the time of retting and the quantity of
water needed. After retting, fibres are collected and laid out
on bamboo racks to dry for 23 days. The weight per unit
length of individual fibres varies from 0.7 to 5.5 tex but the
average is between 1.9 and 2.2 tex. There is no clearly
defined average fibre length and any sample of jute fibres
contains large numbers of short fibres and few long ones.
The approximate chemical composition of jute fibre [9] in
wt.% is: cellulose 71.5, hemicellulose 13.4, pectin 0.2,
lignin 13.1, water soluble compounds 1.2, fat and waxes 0.6.
2.2. Jute yarn
In the present research, hessian yarns which are made
from long jute fibres are used. The yarn is obtained by a
spinning process which depends upon the class of goods
being made; however, there are features common to all
systems. Jute must be softened and lubricated with batch
oil so that the fibre may be processed without excessive
fibre breakage and waste. The messy nature of the reeds

must be split-up and fibres separated as far as possible.


Fibres are drawn evenly into slivers or loose untwisted
strands and then drawn out to the desired yarn thickness.
At the spinning frame the material is given its final drafting
down to the required weight and the fibres are twisted
together to form yarn (245302 tex); this is then wound
on to bobbins. Twisting is done by flyers rotating at speeds
of 35504000 rpm.
2.3. Jute fabric
In this research, a unidirectional type of fabric weave
having a count of 20 12 (for yarns of 245302 tex) is
investigated. 20 12 indicates 20 in number larger yarns
in the warp direction and 12 in number smaller yarns in the
weft direction per inch are used.
3. Fabrication method
A simple hand lay-up technique has been used for
preparation of specimens. The working surfaces are treated
with polyvinyl alcohol (PVA) to facilitate easy removal of
moulds. The matrix material is prepared from commercially
available general purpose polyester resin, accelerator and
catalyst in a weight ratio of 1:0.02:0.026, respectively.
Each layer of fabric is pre-impregnated with matrix material
and placed one over the other in the mould, taking care to
maintain practically achievable tolerances on fabric alignment. Arrangements were made to avoid leakage of matrix
material by keeping the two opposite ends open to allow hot
air to escape during curing. The casting is cured under light
pressure for 30 min before removal of the mould. A fibre
volume fraction of 45 ^ 1% is achieved by using the hand
lay-up technique.
4. Preparation of specimens
All specimens are fabricated individually to avoid voids
and to minimise edge and cutting effects during machining.
Specimens are prepared to required dimensions according to
ASTM-D standards as shown in Table 1.
5. Mechanical testing of fibres, fabrics, polyester resin
and laminates
All mechanical tests except impact tests have been
carried out on a computer-controlled closed-loop servohydraulic MTS-810, having a maximum capacity of 10 metric
ton. For the tests, the 1 metric ton range is used. Load,
displacement and linear strain are measured using a load
cell, linearly variable differential transformer (LVDT) and
clip-on type extensometer, respectively. The longitudinal
and lateral strains used to calculate the Poissons ratio are
measured by strain gauges mounted on the specimen in
conjunction with strain indicator B and K type 1526. PC

T. Munikenche Gowda et al. / Composites: Part A 30 (1999) 277284

279

Table 1
Standards, dimensions and configuration of test specimens
Sample
No.

Specimens tested

ASTM-D Cross head


standard speed (mm/min)

Length Width
(mm)
(mm)

Tabs/span
Depth Guage
Configuration
length (mm) (mm) length (mm)

Tensile:
Jute strand
Jute fabric
Composite/resin

3039
1.3
1.3
1.3

280
280
250

0.65
50
25.4

200
200
172

Compression

3410

1.3

120

25.4

Flexure

790

5.3

130

25.4

Impact

256

63

10

40

Interlaminar shear strength


(ILSS)

2344

1.3

32.7

10

23.9

Inplane shear strength


(IPS)

3518

1.8

250

based Test Work-II, with application software was


employed for composite testing. Analysis of the output
files was by MAT LAB software for post-processing and
graphics. A minimum of five composite samples were tested
to account for statistical scatter and to arrive at mean values.
All the tests are carried out at room temperature.
5.1. Tensile test
Tests on yarn and fabric (count 20 12) having a gauge
length of 200 mm are carried out at a cross-head speed of
1.3 mm/min; the tensile strength and modulus are thus
obtained. The cross-sectional area of a yarn is assumed to
be circular and with the help of a profile projector, the
diameter of the slightly stretched yarn is measured. The
cross-sectional area of fabric is obtained by multiplying
the number of yarns per inch of fabric and average crosssectional area of each yarn in the fabric.
Tests on polyester resin and composite specimens are
carried out to determine tensile strength, modulus of
elasticity and Poissons ratio.
5.2. Compression test
Tests are carried out on polyester resin and composite
specimens to determine compressive strength, modulus of
elasticity and Poissons ratio. The specimens are end loaded,
leaving an unsupported length as the test section.

25.4

12.7

100

172

.65
.75
3.7
2.4

4.2

10
4.6

4.1

200
200
25
12.7

25

initiated by applying the load perpendicular to the fibre


direction. For composite specimens a three-point bend test
and for resin specimens a four-point bend test are adopted.
The relations for flexural strengths and moduli are s 3FL/
2bd 2 MPa, Eb L 3m/4bd 3 GPa, s FL/bd 2 MPa, and Eb
0:21L3 m=bd3 ; respectively, where F load, N; L span
length, mm; the distance between inner and outer loading
point is L/3 and the load span is also equal to L/3 as per
ASTM-D 790 standard; m slope of the loaddeflection
curve at a linear region.
5.4. Impact test
The instrumented impact tester used in this study was a
TINIUS OLSEN Dynatup model 730, pendulum type. A
10 mm hemispherical head impacts a test specimen at a
constant velocity of 1.06 m/sec at an impact energy of
15.26 J. The load and energy absorbed with respect to
time are recorded.
5.5. Interlaminar shear test (ILSS)
This test is mainly used to determine the delamination
effect of the layers and interlaminar shear strength (SH) of
the composites by using the specimen ratios span/
thickness 5, length/thickness 7, and the formula SH
0.75Pb/bd MPa, where Pb breaking load, N.
5.6. Inplane shear test (IPS)

5.3. Flexural test


The polyester resin and composite specimens have a
length/thickness (L/d) ratio of 32:1. The flexural test is

This test is carried out to determine the ultimate inplane


shear strength and shear modulus of the laminate with a fibre
orientation of ^ 458. The load cell and strain gauges are

280

T. Munikenche Gowda et al. / Composites: Part A 30 (1999) 277284

Fig. 1. Stressstrain response of jute fibre and fabric.

used to determine the shear stress t 12 P/2bd and shear


strain g 12 1 x 2 1 y (longitudinal and transverse strain).
The shear modulus is obtained from the formula
G12 st12 =dg12

5.7. Barcol hardness


The hardness of the composite is determined by use of a
Barcol hardness tester. The indicating dial has 100 divisions. The dimensions of test specimens are 150 150
2.6 mm.

6. Results and discussion


Fig. 1 shows the stressstrain diagram for jute fibre and
jute fabric. Jute fibre exhibits stiffer characteristics as
compared to jute fabric. This is due more to initial stretching
of the fabric than the nature of the fibre. For both fibre and
fabric, the curves are initially horizontal due to the stretching effect caused by removing slack from the system.
Strands in the fabric break at different times as each fibre
can stretch independently and break individually when
reaching their breaking stress. The failure mode is by
progressive breaking of the fibres. It has been observed
that strands in the fabric begin to fail from the centre of
the fabric and propagate width-wise on either side. The
ultimate tensile strength and the tangent modulus of elasticity after initial stretching of the as received jute fibre is
found to be 120 MPa and 3.75 GPa; corresponding values
for jute fabric are 85 MPa and 0.8 GPa. The variation in the
values of strengths may be due to: (i) the assumption that
the cross-sectional area of each yarn is circular, (ii) the

Fig. 2. Tensile stressstrain response of jute-reinforced polyester composites and resin.

difference in the strength of individual fibres in the fabric


arising out of process defects.
The tensile strength of a composite material is mainly
dependent on the strength and modulus of fibres, the
strength and chemical stability of the matrix and the effectiveness of the bonding strength between matrix and fibres
in transferring stress across the interface [10]. Fig. 2 shows
the stressstrain diagram for four jute laminates and a
polyester resin. The initial linear portion of the jute laminate
curves show the elastic behaviour of the composite. The
deviation from linearity is an indication of the beginning
of initial matrix cracking, and the first major change in
slope in the curve is the sign of a major crack in the matrix
or the beginning of fibre failure. The first fibre failure occurs
at a stress level of 26 MPa (approx.) and the corresponding
resin stress is 5.5 MPa for the same amount of strain. The
rest of the drops in the curves are indications of progressive
failure of fibres as the applied load increases, and the end of
the curve represents the ultimate stress which is due to fibre
fracture and may be fibre pull-out. However, the failure
mode exhibits breakage and little pull-out of fibres. The
average values of ultimate tensile strength, initial tangent
modulus and Poissons ratio for these composites are
60 MPa, 7 GPa and 0.257, respectively, and corresponding
values for polyester resin are 12.1 MPa, 1.4 GPa and 0.38
(Table 2). The values obtained by the rule of mixture is
63 MPa and 1.12 GPa. The difference in the initial tangent
modulus obtained from experiment and rule of mixture is
due to initial stretching and the nature of the fibres.
When the laminate is loaded in transverse direction-2, it
is observed that the resistance offered by the yarns in direction-1 is more as compared to resistance offered by the yarns
in direction-2. This is because the number of yarns in the
warp direction is greater than the number of yarns in the

T. Munikenche Gowda et al. / Composites: Part A 30 (1999) 277284

281

Fig. 3. Transverse stressstrain response of jute-reinforced polyester


composites and resin.

Fig. 4. Compressive stressstrain response of jute-reinforced polyester


composites and resin.

weft direction per unit dimension. Fig. 3 shows characteristic curves for five jute laminates and a polyester resin. The
values of transverse ultimate tensile strength and transverse
modulus of elasticity are 35 MPa and 3.5 GPa, respectively
(Table 2). In the case of longitudinal tensile tests, the values
for tensile strength and modulus of composite laminate are
almost five times the tensile strength and modulus of the
polyester resin and two times that for the transverse tensile
laminate. The variations in ultimate stress among the same
laminate specimens are due to the inconsistent and highly
non-uniform nature of the jute fibres.
Fig. 4 shows the stressstrain diagram for five identical
jute laminate specimens and one polyester specimen,
subjected to compressive load. The longitudinal ultimate
compressive strength of the composite mainly depends on

the strength of the resin. However, the stiffness (Youngs


modulus) of the composite depends highly on its
reinforcement. This statement is clearly proved and demonstrated in Fig. 4, since Youngs modulus for the composite
and resin is 2.1 GPa and 0.94 GPa, respectively (Table 2). In
general, composite specimens under compression fail by
shear crippling or kinking [11] at the load introduction
point and by a combination of shear and compression. In
the present experiments, specimens are mostly the subject of
a shearcompression type of failure mode. When fibre
buckles, the matrix/fibre interface may fracture in shear
and lead to ultimate failure. Matrix or fibre failure may
begin after a stress level of 22 MPa. The portion of the
curve after the ultimate stress is an indication of progressive
failure of fibres. The average values of ultimate compressive

Table 2
Mechanical properties of jute, polyester resin and jute-reinforced polyester composite

Properties tested
3

Density g/cm
Tensile
a. Jute strand
b. Jute fabric
c. Longitudinal
d. Transverse
Compression
Flexure
Impact KJ/m 2
Inplane shear (IPS)
Inter laminar shear
strength (ILSS)
Barcol hardness
a

Jute laminate (volume fibre fraction Vf 45%)


Strength (MPa)
Modulus (GPa)
Poissons Ratio
1.18(0.15)

Polyester resin
Strength (MPa)

Modulus (GPa)

Poissons Ratio

1.22(0.03)

120(12)
85(3.30)
60(2.80)
35(3.27)
45(2.26)
92.5(5.79)
29(3.30)
16.5(1.05)
10(0.60)

3.75(0.43)
0.8(0.10)
7.0(1.12)
3.5(0.39)
2.1(0.47)
5.1(0.44)

2.2(0.91)

0.25(0.06)
0.22(0.05)
0.392(0.17)

12.1(2.28)

1.4(0.43)

0.38(0.05)

47.1(9.12)
48.0(8.27)
1.76(0.96)

0.94(0.14)
2.2(0.24)

18(4.06)

29.4(0.93)

Numbers in parentheses are standard deviations.

282

T. Munikenche Gowda et al. / Composites: Part A 30 (1999) 277284

Fig. 6. Loadenergytime response of jute-reinforced polyester composite.

Fig. 5. Flexural loaddisplacement response of jute-reinforced polyester


composites and resin.

strength and modulus for the composite are 45 MPa and


2.1 GPa, respectively, and the corresponding values for
the polyester resin are 47.1 MPa and 0.94 GPa (Table 2).
Fig. 5 shows flexural loaddisplacement characteristic
curves for three composites and a polyester resin. Flexural
strength and modulus are controlled by the strength of the
extreme layer of reinforcement. No specimen has failed by
delamination during loading and the failure mode shows
little or no fibre pull-out. It is not possible from loaddisplacement curves to locate the exact beginning of fibre failure
of composite laminates because of the non-linear behaviour
of the jute laminate under flexural load. The crack always
initiates on the tension side of the beam and slowly propagates in an upward direction. Normally, the modulus is very
sensitive to the matrix properties and matrix/fibre interfacial
bonding [12]. Average values for the ultimate flexural
strength and flexural modulus for the composites are
92.5 MPa and 5.1 GPa, respectively, and the corresponding
values for the polyester resin are 48 MPa and 2.2 GPa.
It can be seen from Table 2 that the flexural strength of
the composites at failure is greater than the tensile or
compressive strengths.
The Charpy impact test provides a record of the impact
event. Both loadtime and energytime traces can be
acquired and recorded by an IBM PC/AT during the impact
event. The impact loadtime and energytime traces of a
composite are shown in Fig. 6. None of the composite specimens were broken into two pieces at the middle and all
specimens were pushed through the Charpy anvils during
the impact process. For reinforced specimens like glass/
epoxy [13] and Kevlar/epoxy [14], it has been well documented that the impact force versus time trace is approximately linear to the point of peak force and then drops

suddenly. However, it can be seen from the curve in Fig. 6


that it is linear up to some value and then follows a nonlinear path up to the peak load before dropping suddenly.
Deviation of the loadsignal curve from linearity is an indication of incipient damage. Beaumount et al. [15] termed
the energy to peak load as the initiation energy (EI), and the
energy absorbed after peak load as the propagation energy
(Ep). The ratio of Ep/EI is called the ductility index. Thus, a
low ductility index is supposedly an indication of a brittle
material. The initial jagged load line before the peak value is
reached is believed to be primarily due to stress-wave
propagation and inertial effects such as vibration inside
the strain-gauge tup [16]. These effects are a function of
the relative material and structural stiffness between the
specimen and machine, as well as of instrumentation sensitivity. The average impact energy unit area of composite

Fig. 7. Loaddisplacement response (ILSS) of jute-reinforced polyester


composites.

T. Munikenche Gowda et al. / Composites: Part A 30 (1999) 277284

283

test for four specimens. The curve is linear up to 15 MPa


(approximately) and then follows a non-linear path. A
significant change in the slope of the stressstrain curve
indicates shear failure. Failure of matrix/fibre usually starts
after a stress level of 15 MPa. The composite has mainly
failed due to matrix cracking, fibre failure and debonding of
the matrix/fibre interface. The inplane shear strength and
shear modulus are 16.5 MPa and 2.2 GPa, respectively
(Table 2). Interlaminar shear strength is low compared to
inplane shear strength since fibres also take a part in the
resistance to inplane shear stresses whereas interlaminar
shear is resisted by the matrix.
The Barcol hardness of jute-reinforced polymer composites is found to range between 1525 against 2730 for
polyester resin. The variation in the hardness reading is
caused by the difference in hardness between resin and filler
materials. Addition of jute to the matrix reduces the hardness of the binding material.
Fig. 8. Stressstrain response (IPS) for the ^ 458 tensile test of jutereinforced polyester composites.

and resin specimens are 29 kJ/m 2 and 1.76 kJ/m 2, respectively (Table 2).
For interlaminar shear, the mode of failure is strongly
dependent on the span to depth ratio. All specimens
were simply supported in a fixture and loaded at midspan. Fig. 7 shows the plot of applied load versus
displacement curves for ten identical specimens. It is
well known and also documented, that short beams
usually fail by shear and long beams by tension or
compression [17]. The interlaminar shear strength is
the strength through the thickness of the plies in a
plane containing the fibre axis. No delamination of
plies has been observed during the test. There is a
wide deviation among the curves even though the specimens are all identical. Failure of specimens occurs due
to bending, simultaneous breaking of fibres and partial
pull-out of fibres. It may also be due to weak interfacial
bonding of the matrix or a lack of proper penetration of
resin into the fibres. It is, therefore, very difficult to
explain this behaviour since large numbers of parameters like the breaking of fibres, partial pull-out of
fibres etc., are involved. However, the average interlaminar shear strength of the composites is 10 MPa.
It has been found experimentally for T-300 carbon/epoxy
material that the woven-fabric laminates of ^ 458 inplane
shear test specimens give a better shear response [18]
because of the interlacing of strands and also because of
the higher transverse strength of woven-fabric composites
as compared to unidirectional composites. The effect of
transverse stress on shear failure is reduced considerably.
In the case of the ^ 458 tension test, even though normal
stresses are present the failure is essentially due to shear.
Fig. 8 shows the stressstrain diagram for the ^ 458 tensile

7. Conclusion
It is concluded that although the mechanical properties of
jute/polyester composites do not possess strengths and
moduli as high as those of conventional composites, they
do have better strengths than wood composites [19] and
some plastics. Therefore, these composites could be considered for future materials use. Since the reinforcing material
is eco-friendly, non-toxic, non-health hazardous, low in cost
and easily available as compared to conventional fibres like
glass, Kevlar, asbestos etc., the composites are a good
substitute for wood in indoor applications such as shelves,
partitions, wash basins [20] and table tops, and may also be
suitable for outdoor uses such as roofing, drainage pipes,
automobile components, electrical fittings as well as larger
items such as lightweight fishing boats [21]. To ascertain
their suitability for outdoor applications, a few more tests
are to be carried out to evolve the hygrothermal and weather
resistance properties of these composites.

Acknowledgements
The authors gratefully acknowledge the financial
support provided by the All India Council for Technical
Education, New Delhi, and the B.M. Srinivasaiah
College of Engineering, Bangalore. We are sincerely
thankful to Dr K.G. Narayanan, Director, and the staff
of the Structures and Materials division, in particular,
G.S. Ravindra, Scientist D and V.P. Rangaiah of the
Aeronautical Development Establishment, Bangalore,
for providing us with testing facilities. The first author
is grateful to his parent organisation, the S.J.C. Institute
of Technology, Chickballapur, for deputing him to
higher studies.

284

T. Munikenche Gowda et al. / Composites: Part A 30 (1999) 277284

References
[1] Shah AN, Lakkad SC. Mechanical properties of jute reinforced plastics. Fibre Science and Technology 1981;15:4146.
[2] Bhattacharyya DN, Chakravarti IB, Sengupta SR. J. Sci. Industries
Res. 1961;20D:193.
[3] Verma IK, Anantha Krishnan SR, Krishna Murthy S. Composites of
glass/modified jute fabric and unsaturated polyester resin. Composites
1989;20:383388.
[4] Mohan R, Kishore, Shridhar, Rao RMVGK. Compressive strength of
juteglass hybrid fibre composites. Journal of Material Science
Letters 1983;2:99102.
[5] Chawla KK, Bastos AC. The mechanical properties of jute fibre and
polyester/jute composite. Mechanical Behaviour of Materials
1979;3:191196.
[6] Winfield AG. Proceedings of second symposium on new fibres and
composites, sponsored jointly by Dept. of Science and Technology,
India, and UNIDO, January 1011, 1977:31.
[7] Winfield AG. Plastics and Rubber International 1979;4:23.
[8] Atkinson RR. Jute Fibre to Yarn. Bombay: B.I. Publications,
1965:2737.
[9] Carlos A Cruz-Ramos. Natural Fibre Reinforced Thermoplastics,
Mechanical Properties of Reinforced Thermoplastics. Amsterdam:
Elsevier, 1986: 6581.
[10] Gao Z, Reinfshinder Kenneth L. Tensile failure of composites, influence of interface and matrix yielding. Journal of Composites Technology and Research 1992;14:201202.
[11] Carlsson, Leif A, Pipes R Byron. Experimental Characterisation of
Advanced Composite Materials. New York:Prentice-Hall, 1987:69
81.
[12] Jang BJ, Liau JY, Hwang LR, Shih WK. Particulate and whisker

[13]

[14]

[15]
[16]

[17]

[18]

[19]
[20]

[21]

modification of matrix resin for improved toughness of fibrous


composites. Journal of Reinforced Plastics and Composites
1990;9:314332.
Agarwal Bhagvan D, Broutman Lawrence J. Analysis and Performance of Fibre Composites. New York: Wiley Interscience,
1980:249.
Adams DF. Impact response of polymer matrix composite materials.
Composite materials: Testing and Design (Fourth Conference).
ASTM STP 1976;617:418419.
Beaumount PWR, Riewald PG, Zweben C. Foreign object impact
damage to composites. ASTM STP 1975;568:134158.
Helfinstine JD. Charpy impact of unidirectional graphite/aramid/
epoxy hybrid composites. Composite Materials: Testing and Design
(Fourth Conference). ASTM STP 1976;617:375388.
Zweben C, Smith WS, Wardle MM. Test methods for fibre tensile
strength composite flexural modulus and properties of fabric-reinforced laminates. Composite Materials: Testing and Design (Fifth
Conference). ASTM STP 1979;674:262288.
Naik NK, Ganesh VK. Failure behaviour of plain weave fabric laminates under inplane shear loading. Journal of Composite Technology
and Research 1994;16(1):15.
Bodig Jozsef, Benjamin A Jayne. Mechanics of Wood and Wood
Composites. New York: Van Nostrand Reinhold, 1982:282.
Satyanarayana KG, Kulkarni AG, Sukumaran K, Pillai SGK, Cherian
KA, Rohatgi PK. On the possibility of using natural fibre composites.
In: Composite Structures. Amsterdam: Elsevier, 1981:618.
Ajit Bantia K, Subal C Bera, Banerjee. Jute/shan based natural fibre
composites. The first national symposium on recent trends in the
development of the composite materials, December 1983, Pune,
India, 1983:51.

You might also like