Вы находитесь на странице: 1из 59

Ground Improvement by Alkali Activated

Binder Treated Jute Geotextile

Undergraduate Thesis

Submitted in partial fulfillment of the requirements of


BITS F421T Thesis

By

Shashank Gupta
ID No. 2014A2TS0584H

Under the supervision of:

Dr. Anasua Guharay


&
Dr. Arkamitra Kar

BIRLA INSTITUTE OF TECHNOLOGY AND SCIENCE PILANI,


HYDERABAD CAMPUS
December 2017
Certificate
This is to certify that the thesis entitled, “ Ground Improvement by Alkali Activated Binder
Treated Jute Geotextile” and submitted by Shashank Gupta ID No. 2014A2TS0584H in partial
fulfillment of the requirements of BITS F421T Thesis embodies the work done by him under my
supervision.

Supervisor Co-Supervisor
Dr. Anasua Guharay Dr. Arkamitra Kar
Assistant Professor, Assistant Professor,
BITS-Pilani, Hyderabad Campus BITS-Pilani, Hyderabad Campus
Date: Date:

i
“If four things are followed - having a great aim, acquiring knowledge, hard work, and perseverance
- then anything can be achieved.”

A. P. J. Abdul Kalam
BIRLA INSTITUTE OF TECHNOLOGY AND SCIENCE PILANI,
HYDERABAD CAMPUS

Abstract
Bachelor of Engineering (Hons.)

Ground Improvement by Alkali Activated Binder Treated Jute Geotextile

by Shashank Gupta

Geotextiles are popular for reinforcing soil, improving drainage, controlling soil erosion, and
embankment construction. Existing research recommends the improvement of soil in an economic
and eco-friendly manner by using jute geotextiles. However, jute fibers have the tendency
to degrade in the acidic and alkaline environment of the soil. Jute geotextile treated with
antimicrobial chemicals are used as a substitute for manmade geosynthetics as it improves
the life expectancy of jute. But these chemicals are expensive and are a potential source of
leaching. Besides, there is increasing interest in the application of alkali-activated binders (AAB)
in engineering practices. Thus, in this study, an intensive chemical study has been done and
is correlated with the mechanical properties of treated jute. .Moreover, engineering properties
of jute treated with AAB are compared with those of untreated jute. Experimental studies on
vertical permeability, shear strength properties and bearing capacity of sand reinforced with
treated and untreated jute are also conducted. These studies show that the load bearing capacity
increases by 35% approximately and the vertical permeability of sand do not vary significantly
with the inclusion of AAB treated jute. However, a notable decrease in friction angle of 38.3%
between reinforcement and sand is observed due to the treatment. Thus the technique proposed
in this study has the potential for implementation in practical applications.
Acknowledgements
I wish to express my sincere gratitude to my supervisor, Dr. Anasua Guharay.I would like to
thank her for providing me laboratory facilities along with proper guidance and encouragement
at each and every stage of project. Her knowledge in the field of the experimental geomechanics
played a pivotal role in my project. and I am very fortunate to work with Dr. Guharay on this
topic.

My sincere appreciation goes to Dr. Arkamitra Kar, who taught and helped me to complete the
chemical characterization of treated jute.Without his extensive and remarkable knowledge on
the experimental chemistry, the project will not be completed. I am extremely grateful to work
with Dr. Kar.

I would like to convey my gratitude to Dr. Sridhar Raju and Dr. Murari Varma for agreeing to
be part of my evaluation committee.

I would like to thanks Mr. V. Purneshwar for helping me in several lab experiments and for
providing positive feedbacks at each and every stage of my work.

I would like to thanks lab technicians, Mr. Hemanth Mudhiraj and Mr. M. Basha, to help me
carrying out several experiments.

iv
Contents

Certificate i

Abstract iii

Acknowledgements iv

Contents v

List of Figures vii

List of Tables viii

Abbreviations ix

1 Introduction 1
1.1 Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.3 Research Objectives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3

2 Literature Review 4
2.1 Application of JGT . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
2.2 Jute Treatment Processes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
2.2.1 Scouring and Bleaching . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
2.2.2 Rot Resistant Solution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
2.2.3 Acetylation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
2.2.4 Transesterification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
2.3 Jute Composites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9

3 Materials and Treatment Solution 12


3.1 Materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
3.2 Treatment Solution Preparation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13

4 Chemical Characterization of Treated Jute 15


4.1 X-Ray Diffraction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
4.2 Fourier Transform Infrared Spectroscopy . . . . . . . . . . . . . . . . . . . . . . . 16

v
Contents vi

4.3 Surface Morphology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18


4.4 Thermogravimetric Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20

5 Mechanical Properties of treated geotextile 26

6 Sand reinforced with treated jute 28


6.1 Vertical Permeability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
6.2 Interface properties of sand and reinforcement . . . . . . . . . . . . . . . . . . . . 28
6.3 Model Plate Load Test . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31

7 Numerical Study 36
7.1 Footing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
7.2 Soil and Geotextile . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
7.3 Mesh . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
7.4 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39

8 Durability Study 40

9 Conclusion 42

10 References 44

A MATLAB Code for Curve Fitting 47


List of Figures

3.1 JGT before and after treatment . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14

4.1 X-Ray Diffractogram for untreated and treated jute sample . . . . . . . . . . . . 17


4.2 FTIR spectral pattern of (a) Untreated jute, (b) 0.35 w/s AAB treated jute, (c)
0.40 w/s AAB treated jute, and (d) 0.45 w/s AAB treated jute. . . . . . . . . . . 19
4.3 SEM images of untreated jute at X2000 and X5000 magnification along with EDS
analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
4.4 SEM images of 0.35 AAB treated jute at X2000 and X5000 magnification along
with EDS analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
4.5 SEM images of 0.40 AAB treated jute at X2000 and X5000 magnification along
with EDS analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
4.6 SEM images of 0.45 AAB treated jute at X2000 and X5000 magnification along
with EDS analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
4.7 Thermogravimetric curves of untreated and AAB treated jute . . . . . . . . . . . 25

5.1 Tensile strength of raw jute and treated jute geotextile . . . . . . . . . . . . . . . 26

6.1 Permeability of reinforced sand . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29


6.2 Friction angle between sand and reinforcement . . . . . . . . . . . . . . . . . . . 30
6.3 Cohesion between sand and reinforcement . . . . . . . . . . . . . . . . . . . . . . 30
6.4 Interface Fricion angle vs Shear Strain Rate . . . . . . . . . . . . . . . . . . . . 31
6.5 Experimental Setup of model plate load . . . . . . . . . . . . . . . . . . . . . . . 32
6.6 Load vs Displacement graph at the depth of 10 cm . . . . . . . . . . . . . . . . . 33
6.7 Load vs Displacement graph at the depth of 15 cm . . . . . . . . . . . . . . . . . 33
6.8 Load vs Displacement graph at the depth of 20 cm . . . . . . . . . . . . . . . . . 34
6.9 Load vs Displacement graph at the depth of 25 cm . . . . . . . . . . . . . . . . . 34

7.1 Meshing with (a) Soil (b) Soil reinforced with geotextile . . . . . . . . . . . . . . 37
7.2 Deformed mesh for (a) Soil, (b) Soil reinforced with AAB treated jute geotextile. 38
7.3 Comparison of load-settlement curves from numerical and experimental study. . . 38

8.1 Treated and untreated jute exposed to unsterile soil (1 month) . . . . . . . . . . 40


8.2 Tensile strength of geotextile after exposure to unsterile soil . . . . . . . . . . . . 41

vii
List of Tables

3.1 Properties of sand . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12


3.2 Amount of AAB required treating per sq. metre jute geotextile . . . . . . . . . . 13

4.1 Bonds corresponding to characteristic peaks . . . . . . . . . . . . . . . . . . . . 16

viii
Abbreviations

AAB Alkali Activated Binder


CBR California Bearing Ratio
EDS Energy Dispersive X-ray Spectroscopy
FTIR Fourier Transform Infrared Spectroscopy
JGT Jute Geotextile
LVDT Linear Variable Differential Transformer
SEM Scanning Electron Microscope
TJP Treated Jute Polypropylene composite
UJP Untreated Jute Polypropylene composite
XRD X- Ray Diffraction

ix
I dedicate this work to my father, who never lost confidence in me.

x
Chapter 1

Introduction

The present study primarily aimed at determining the mechanical and chemical properties of
jute geotextile treated with alkali activated binder and further study the interaction of the
treated jute geotextile with sandy soil. A brief study regarding the durability of treated jute is
also presented. Moreover, few numerical studies have been done to bouldered the experimental
results. This chapter presents the background information of this treated geotextile along with
the motivation for the study. The chapter concludes with research objectives.

1.1 Background

The ground improvement is one of the problems that have been troubling humankind since the
genesis of the civilisation. In past, the Roman civilisation improved the ground strength by
mixing straw with infill soil and early Asian civilisation used elephants for the compaction of soil
(Rujikiatkamjorn et al., 2005). Natural geotextiles were also used in ancient constructions for the
purpose of ground improvement. The mats made up of reeds and tree twigs were used in temples
of Babylonia and Great Wall of China, respectively (Sanyal, 2017). In 1953, the geosynthetics
were first used on a large scale in the Delta Project that included series of constructions in
Netherlands as protection against floods. Two decades later, the extruded geosynthetics were
used in construction in England and Japan. By eighties, the high-density polyethene (HDPE)
appeared in markets and soon become a widespread choice as a drainage medium ( Maione et al.,
2000). After the eighties, the geosynthetics drew a lot of popularity as a filter, a soil reinforcer,
and a separator. The consumption of geosynthetics has grown from 10.2 million sq. meters to
colossal 2475 million sq. meters in last three decades. However, the natural geotextiles like jute,
coir, hemp etc. constitute only 5-6% of total geotextile consumption (Sanyal,2017).

On the basis of mechanical properties, natural geotextiles are equivalent to synthetic geotextiles.
The natural fibres have the similar specific tensile strength and Young’s modulus as compared
1
Chapter 1. Introduction 2

to synthetic fibres (Nam and Netravali, 2006). Nevertheless, as pointed out in above paragraph,
the global use of natural geotextile is considerably low. The possible reasons for the limited use
of natural geotextiles can be the (i) Paucity of academic research on the properties of natural
geotextile along with the behaviour and performance of soil reinforced with them, (ii) irregularity
in properties between the same kind of geotextile due to different origin and manufacturing
processes, and (iii) low resistance to biological, physical and weathering degradation.

However, the study in the field of natural geotextile has been paced up in the recent years. The
vast laboratory studies and more than 260 field trials were conducted till date in India (Sanyal,
2017). The literature review also shows that researchers have developed the treatment procedures
to improve the mechanical properties and durability of jute. An extensive study is done on the
chemical changes due to treatment by X-Ray Diffraction (XRD), Fourier Transform Infrared
Spectroscopy, Scanning Electron Microscope (SEM), Thermal Gravimetric Analysis(TGA),
Chemical Composition Analysis etc.(Wang et al., 2009 and Saha et al., 2012 ) Furthermore, they
have correlated these chemical changes to durability and mechanical properties of jute geotextile
(Saha et al., 2012 and Anderrson and Tillman, 1989). The researchers had also conducted a series
of experiments to determine drainage, bearing capacity, and shear properties of soil reinforced
with geotextile(Li et al., 2014 and Ahmad et al., 2010). However, an insignificant amount of
work has been found regarding the standardization of the use of natural geotextile.

1.2 Motivation

Most of the present market is based on the manmade geotextiles like polypropylene and HDPE.
However, with deteriorating environment, natural alternatives are required. The production of
synthetic geotextile consumes high energy than natural geotextiles. Moreover, natural geotextile
doesn’t leave CO2 footprint and cost less in contrast to synthetic geotextiles (Wambua et al.,
2003). In addition, there are several construction applications where natural geotextiles can
be opted instead of synthetic one. For instance, river banks or hill slopes requires protection
until the vegetation is grown enough to support the ground and control soil erosion. The time
for support may vary with location, climate, soil type, and species of vegetation. In India, raw
jute geotextile (JGT) and treated JGT has been implemented for numerous applications. Datta
(2007) has presented various case studies in which JGT have been used to control soil erosion,
improve subgrade for road construction, and protect river and canal bank. However, the life can
be too short in some conditions and hence it may degrade before fulfilling its tasks. Therefore a
proper treatment procedure is required to enhance the properties of jute by not incurring too
much cost.
Chapter 1. Introduction 3

1.3 Research Objectives

The present study holds the following objectives:

• To discuss the existing treatment procedures briefly.

• To treat jute with alkali activated binder(AAB) of different water to solid (w/s) ratios and
standardized the amount of required raw materials to treat per sq meter of jute fabric.

• To study the chemical and microstructural characterization of jute fibre due to AAB
solutions.

• To study the drainage properties, shear properties, and bearing capacity of sand reinforced
with AAB treated JGT.

• To validate the plate load results numerically.

• To compare the durability of untreated JGT and AAB treated JGT.


Chapter 2

Literature Review

This chapter presents few case studies to show the use of the JGT in various construction appli-
cations. The existing treatment procedures of jute, with their pros and cons, are also presented
in the chapter.In addition, the recent studies on the jute-manmade synthetics composites have
been discussed.

2.1 Application of JGT

JGTs have been used for stabilising the bank slopes by reducing soil erosion. Sanyal and
Chakraborty (1994) have presented a study in which the slope of Nayachara island, situated in
the middle of Hooghly river, was stabilised by bitumen treated jute geotextile.

The western bank was protected with a layer of treated jute geotextile covered with riprap of
different sizes. The slope of the bank was kept as 1 to 5. An anchor and a toe beam were
constructed at the top and the bottom of the slope, respectively, to hold jute geotextile. After
1.5 years, no disturbance or damage was observed in the geotextile protected part. However, the
tensile strength of geotextile was dropped by 70% due to biological and physical degradation
under riprap.

Jual and Sharda (2008) have presented a study in which JGTs were used in hilly areas of the
states of Himachal Pradesh and Uttarakhand. These sites were, at that time, damaged by
various landslides and mining activities. A single layer of open weave Jute geotextile of 500 gsm
(grams per square meter) was used to reinforce the damaged area. Grass and horticulture plants
were also planted in the reinforced soil.

After 3 years, rich vegetation was established in that region. Also, the untreated area was
protected by the treated region. The treatment was done on top of hilly terrain so at rainfall the

4
Chapter 2. Literature Review 5

speed of flowing water decreases and erosion at untreated area decreased. The authors estimated
that twice the region of treatment was protected by treated area and this region was suitable for
farming too.

As jute Geotextiles is biodegradable, it degraded after some time, but that time was enough for
vegetation to grow. This vegetation provided strong resistance against the erosion of soil. In
this case study, 44000 ha of hill roads and 25000 ha of mining area were treated. The authors
also estimated the cost of treatment to be Rs27/m2.

Rao(2003) have presented a case study in which a 360m length of the unpaved road was
constructed near Kakinada Port in the Andhra Pradesh state of India. A single layer of
untreated Jute geotextile of 750gsm was used in the construction. The top strata of soil profile
consisted of soft clay, underlain by sandy silt soil. The soaked and unsoaked CBR values of
soil were 1.61% and 2.1% respectively. After compaction of the top layer of soil, a 3mm Jute
geotextile was laid. Above this compacted layer of sand and gravel were laid.

Data corresponding to water content, dry density and void ratio were collected after regular
interval of time. It was observed that the water content of soil decreased with time. There was a
drop of 13% and 35% in the water content of soil after 3 months and 21 months, respectively,
of construction. On the other hand, dry density of soil increased with time. After 3 months
of construction, dry density increased by 21%. The void ratio decreased with time, but not
significantly.

CBR test was performed on soil after 30 months of construction. It was found that the unsoaked
CBR and soaked CBR were 2.87 and 2.96 times, respectively, higher than initial Values of CBR
(mentioned above). The road was visually inspected after 17 years, and there was no visible
damage to roads.

2.2 Jute Treatment Processes

The researchers have treated jute with several chemicals to improve it’s strength and durability
properties. These treatments either yield chemical changes or physical changes to improve the
strength of jute geotextile. In the following sections, the efficacy, treatment processes, and
shortcomings of few treatment processes are discussed in detail.

2.2.1 Scouring and Bleaching

Wang et al. (2009) treated raw jute with two separate chemical processes, alkali scouring and
hydrogen peroxide bleaching. They treated jute with a dilute alkali solution and hydrogen
Chapter 2. Literature Review 6

peroxide solution for 90 min at 95◦ C and 90◦ C respectively. The jute to liquid ratio was kept
1:20 for both treatments. After each treatment, samples were neutralized with dilute sulfuric
acid.

The tests conducted on prepared treated samples were FTIR analysis, surface characteristic
analysis, the chemical composition of fibre, XRD analysis, Breaking tenacity, breaking elongation,
fineness index test, moisture regain test and optical properties test. Brief procedure and results
obtained are discussed as follows. FTIR analysis showed that hemicellulose and lignin are
partially removed by both scouring and Bleaching. Although Alkali scouring was comparatively
better than bleaching in the removal of non-cellulose material. The chemical composition of fibre
shows similar results. The percentage of cellulose rose from 62% to approx. 78% for scoured
sample and 76% for the bleached sample. It was also observed that scouring was better than
bleaching in removing Hemicellulose while bleaching was better at removing Lignin. Another
sample prepared by scouring followed by bleaching had the highest percentage of cellulose ( 82%).

SEM analysis of raw jute shows that there were many impurities like oils, fats, waxes, lignin
etc. After performing SEM analysis on a treated sample, it was observed that alkali scouring is
better than bleaching in removing surface impurities. The sample that was both scoured, and
bleached, had the cleanest surface than other samples. X-ray diffraction analysis show there was
no crystalline transformation i.e. there was no change in the structure of crystals. But there was
a change in the percentage of various crystals present in raw jute hence there was variation in
crystallinity among various samples. It increased for all treated sample but to different extents.

Crystallinity was highest for Bleached and scoured sample followed by the scoured sample.
The bleached sample had the lowest crystallinity among treated samples. Various tests were
conducted to obtain the Physio- mechanical properties of treated samples. As there was an
increase in crystallinity and decrease in the amount of hemicellulose and lignin from jute sample,
breaking tenacity increased for all samples. The scoured sample had highest breaking tenacity
followed by bleached sample, further followed by scoured and bleached sample. For breaking
extensions results obtained were not expected. There was an increase in breaking extension for
the scoured sample, but for bleached and “scoured and bleached” samples, breaking extension
decreased. Fineness Index of all treated samples decreased due to the removal of lignin from jute.
Moisture regains value decreased but not significantly. The removal of hemicellulose and lignin
decreases moisture regain but on same hand removal of fat, oil and waxes increase moisture
regain. So both factors cancel out each other. Optical Properties test show that brightness and
whiteness increased for all treated samples. Yellowness index increased for Scoured sample and
decreased for the bleached sample.
Chapter 2. Literature Review 7

2.2.2 Rot Resistant Solution

Prodhan (2008) emphasized the importance and advantage of chemically treated Jute. The author
also mentioned various treatment procedures to enhance the life of jute geotextile. According to
the author, jute life is short under the soil as it starts degrading after six months of placement.
Jute is also sensitive to both acidic and alkaline solutions. Prodhan proposed series of chemical
treatments that would likely extend the life of jute to 5-10 years under the soil. Firstly, he
proposed to spray jute with a solution of sulphate of copper, sodium carbonate and water. This
treatment increases the rot resistance of jute. Further, he treated it with an emulsion prepared
from a volatile oil and bitumen. This emulsion was applied to jute surface with a brush. Then
he left the treated sample under sunlight until it dried off. The sample was further treated with
silicate solution and dried again. Finally, Ca-based grease mixed with carbon black and volatile
oil was rubbed on modified jute cloth. This treated sample was placed under a 500mm layer of
the subgrade.

2.2.3 Acetylation

Anderrson et al.( 1989) simply treated the jute with acetic anhydride. These treated samples
were tested for tensile strength, rot resistance and hydrophobicity. Authors dipped a 1*1m2
Jute cloth in acetic anhydride for a minute. Then they heated the dipped sample at 90◦ C and
120◦ C for 5min.-24hr. Excess acetic anhydride and acetic acid, a by-product, were removed
after heating. In some tests, the author added the different amount of acetic acid in anhydride
solution. After the treatment, the weight gain due to acetylation was noted down for various
samples. The weight gain due to acetylation increased with reaction time, although its rate
of increment decreases after 2 hours. Samples heated at 120◦ C had a higher weight gain than
samples heated at 90◦ C. This show that acetylation of jute is an endothermic reaction. It was
noted that weight gain after 2 hours at 120◦ C and 90◦ C was 11% and 7.5% respectively. Addition
of acetic acid increases the weight gain of jute as it has a swelling effect on jute that helps it to
react easily with acetic anhydride. Although an increase in concentration resulted in dilution of
acetic anhydride hence decreased the weight gain percentage.

To check the resistance against rotting, both untreated and treated jute samples were kept in a
fungus cellar in moist soil at 25◦ C. This cellar had various fungi and bacteria in it. Samples
were visually inspected for damage after 2, 5.5 and 8 months. The author concluded that sample
with 11.5% weight gain was not damaged at all after 2 months. Even after 5.5 months, it was
moderately damaged. Although after 8 months there were signs of severe damage. Samples with
14.2% and 16.2% weight gain showed better resistance against rotting as even after 8 months
there were not any signs of significant damage. Untreated jute samples under the same condition
were severely damaged after 2 months and after 5.5 months they were disintegrating. Treated
Chapter 2. Literature Review 8

samples were also soaked in water. The weight increase due to water absorption was noted down.
It was found that the weight increase due to water uptake was reduced. For samples that gain
16% weight during acetylation, water absorption was reduced by 30% compared to untreated jute
sample. Also, the treated jute samples shrink less than untreated samples. Breaking strength of
jute cloth decreased after acetylation. In weft direction, there was a reduction of about 16% and
42% in the strength of samples with 11 % and 16 % weight gain respectively. In warp direction,
the corresponding reductions in strength were about 10% and 33%. The author concluded that
sample with 11% weight gain is the best sample as there was a significant increase in the life of
jute at the cost of little loss of strength.

2.2.4 Transesterification

Saha et al. (2012) treated the jute sample with a chemical mixture mostly obtained from
natural sources. The chemical mixture consisted of sodium hydroxide, cashew nut shell liquid,
plant tannin, neem oil, resorcinol and formaldehyde in 1:8:10:6:2:4 proportion by weight. They
maintained the pH of 8 of the solution. Geotextile was dipped for 24 hrs and then was heated
in an oven for 1hr. They prepared two samples, one by treating jute fibres (they wove the
fabric with treated fibre) and another by treating jute at fabric level. For the sake of simplicity,
geotextile wove with treated fibre is referred by T1 and the sample treated at the fabric level is
referred as T2. Untreated jute is referred by JU.

Once samples were prepared, tests and analysis were conducted to get an idea of the characteristics
of treated samples. Tensile strength test was performed on freshly prepared samples following
ASTM D751. T1 sample had 13% lower tensile strength than JU sample while T2 had 13%
higher tensile strength than JU sample. Although breaking elongation length did not alter much.
Treated samples were also tested for their resistance to chemical and biological degradation.
After keeping samples in 3% NaCl solution for 120 days, tensile strength test was conducted
on them. It was found that T1 and T2 sample retained 82% and 64% of their initial strength
respectively. On the other hand, JU sample retained only 17% of their initial strength. Treated
samples were also kept in solutions of PH varying from3 to 10. They found that T1 and T2
samples retained 50% of their initial tensile strength in PH 4-9 solutions, although T1 samples
had somewhat better results. JU samples performed poorly in Tensile strength test as they
retained only 15% of their initial strength.

For testing geotextile sample resistance against biological degradation, a simulation of real-life
condition had been tried to set up. Authors made mixture of black soil, cow dung and sand with
the appropriate amount of water. They buried the sample under this mixture for 200 days. The
soil was also replaced with a frequency of 7 days. This setup was kept in controlled environment
w.r.t temperature and humidity. It was found that treated samples had high resistance against
Chapter 2. Literature Review 9

biological degradation. T1 and T2 samples retained 40-50% of initial strength while UJ sample
lost it most of the strength in 90 days due to complete disintegration. The resistance against
physical degradation was observed by keeping the samples in artificial weather chamber for 500
hrs. T1 and T2 samples retained 80% and 75% of their strength respectively, while UJ retained
only 33% of its initial strength. It was also observed that water affinity of T1 and T2 samples
were decreased by 66% and 55% respectively.

To study the chemical changes in Jute, FTIR analysis was done on jute JU, T1 and T2 jute
samples. A sample of JU, T1 and T2 were washed with alcohol to remove impurities. Then
they were hydrolyzed with 1 % NaOH solution. These modified samples were analyzed by FTIR.
FTIR results showed the presence of stearic and oleic acid in T1 and T2 samples. This inferred
the transesterification of jute as these acids are the main constituent of neem oil. FTIR analysis
also showed that hydroxyl group were altered in both T1 and T2 samples, although the T1
sample was more affected. This indicates that jute treated at fibre level was more affected than
jute treated at fabric level. XRD results showed that there was an increase in crystallinity of
jute samples. Crystallinity increased by 9% and 4% for T1 and T2 samples respectively. Saha
et.al concluded the paper with a half-life (the time by which sample has lost 50% of its initial
tensile strength) estimation of T1 and T2 samples. The estimated life of JU, T1 and T2 samples
was 384, 1584 and 1115 days respectively.

2.3 Jute Composites

Basu et.al (2009) prepared woven Jute- HDPE (85%/15%) blended geotextile. Jute was used in
cross direction while HDPE was used in the machine direction. The mechanical properties of
freshly prepared samples were determined. After that, this blended geotextile was used in the
construction of a unpaved road in West Bengal. After 18 months that road was analyzed and
comparison was done between parts of the road where geotextile was used and where it was not
used.

The tensile strength test was conducted on a blended sample in both machine and cross directions.
The results were different in both directions. In machine direction, the average breaking tensile
strength and elongation at break were 10.47kN/m and 22.67% respectively. On the other hand,
in the cross direction, the corresponding values were 15.2kN/m and 7.16%. This result shows
that jute that was used in cross direction has higher tensile strength than HDPE but lower
elongation at break. So, the overall tensile strength of the blended fabric is more than 100%
HDPE sample. Jute-HDPE blended fabric had higher resistance to CBR puncture than pure
100% PP sample. At the CBR probe displacement of 27mm, the force resisted by 100% PP
geotextile was around 1kN, but for the same displacement, Jute- HDPE fabric was resisting
1.5kN force. The author also showed that filtration property of blended geotextile by showing
Chapter 2. Literature Review 10

the relation between apparent opening size and water flow rate. Cost estimation of blended
geotextile was also done by authors. They estimated that per m2 of blended geotextile would
cost 1.32-1.36 dollars that were little costly than a cost of 100% HDPE and little cheaper than
Jute.

Field trial of Jute-HDPE was done on medium traffic road. The soil was silty clay. The region
in which geotextile was used had, high rainfall. During rain, the water level was 1m-2m below
ground level. The water content in soil was also high as a brook flowed near the road. The soil
was properly compacted and 1.5mm thick layer of geotextile was laid on it. Above the geotextile
two layers, subgrade and top layer each of 100 mm were laid. The road was analyzed after 18
months. During this period a season of monsoon was passed with the heaviest rainfall of 216cm.
Visual inspection of the road was done. It was found that there were not any potholes and surface
cracks on the geotextile reinforced road while the part of the road without Geotextiles had no. of
potholes and surface cracks. The rutting in non-treated part 8-43mm while in geotextile treated
part rutting was not even measurable. CBR test was conducted after 11 months and 18 months.
After 11 months, the CBR value of geotextile reinforced road was 67.56% more than a road
without geotextile. This gap increased to 76.74% in 7 months. This result showed that rate of
drop of strength is more in soil without geotextile than that of soil with geotextile.

Rawal and Sayeed (2014) used non-woven needle punched jute/PP fibres as a geotextile. They
treated the jute with 4% wt. NaOH solution for 30 min. and prepared two kinds of hybrid
Geotextiles. One is hybrid of untreated jute and polypropylene fibres (UJP) and another one is
hybrid of treated jute and polypropylene fibres (TJP). They also varied the Jute to PP ratio to
get various samples. After preparing the treated Jute sample, Rawal and Sayeed(2014) compared
the tensile strength properties of treated jute with that of untreated jute and Polypropylene.
The tensile strength of the treated jute sample was 44% higher than untreated jute sample,
although it was more or less equal to the tensile strength of polypropylene. Breaking extension
of the treated jute sample was slightly higher than untreated jute and was about 23 times lower
than that of PP fibre. Also, the modulus was highest for treated jute followed by untreated jute
and PP fibre.

After performing tests on each type of fibres, tensile strength tests were conducted on UJP
samples and TJP samples in both machine and cross directions. Three proportions of Jute and
PP were used 20/80, 40/60 and 60/40. It was observed that tensile strength in cross direction
for each and every sample was higher for TJP sample compared to UJP sample. Although
with increment in the proportion of Jute fibre in hybrid, this gap in strength was seemed to
be reduced. It was also noted that the 40/60 TJP sample had the highest tensile strength of
all samples. Breaking extension was seemed to reduce with an increase in jute proportion in
hybrid fibre. It was observed that 20/80 UJP and 40/60 TJP samples had the highest and lowest
breaking extension respectively. There was not much change in mechanical properties in the
Chapter 2. Literature Review 11

machine direction, although Tensile strength and breaking extension increased by little. Also,
the jute hybrid fibre with Jute more than 40% did not show much improvement. The authors
concluded that best choice for optimal results in the field was 40/60 TJP sample.
Chapter 3

Materials and Treatment Solution

3.1 Materials

Commercially available jute (Corchorus olitorius) hessian cloth was used for this study. The jute
brought was available in a roll of 0.91m (1 yd) width and 30m length. The sodium hydroxide
pellets and sodium silicate solution were used for the preparation of the solution. The purity of
sodium hydroxide pellets was 99%. The sodium silicate solution was composed of 55.9% water,
29.4% SiO2, and 14.7% Na2O.

Locally available river sand, collected from Godavari River basin, Telangana, India is used in
the present study. The sand is classified as poorly graded sand (SP) according to Unified Soil
Classification System. The maximum and minimum dry densities are measured conforming IS
2720: 1983 Part 14. A constant relative density of 40% is maintained throughout the experimental
procedure to stimulate medium dense sand condition. The properties of cohesionless soil used
are presented in Table 3.1.

Table 3.1: Properties of sand

Properties Values
Coefficient of Uniformity (Cu) 2.5
Coefficient of Curvature (Cc) 0.9
Classification SP
Internal Angle of Friction 26.1
Maximum Void Ratio 0.637
Minimum void ratio 0.411

12
Chapter 3. Materials and Treatment solution 13

3.2 Treatment Solution Preparation

The step-wise preparation of treated jute geotextile is described below.

1. The treatment mix consists of Class-F fly ash, sodium silicate, sodium hydroxide and water.
The ratio of fly ash to sodium silicate to sodium hydroxide by mass is 400:129.43:10.57.

2. The first step is to prepare alkaline activator solution. This solution consists of sodium
hydroxide solution, sodium silicate, and water. The water content can be varied to prepare
solutions of different water to solid ratios. However, the amount of water present in sodium
silicate solution has to be deducted while calculating the additional water to be added to attain
the target water content. In this study, three different water-to-solid ratios were used, viz. 0.35,
0.40, and 0.45. The solution can be prepared by blending the constituents and accompanied by
constant stirring till a clear blend is obtained. As the reaction between sodium hydroxide and
sodium silicate solution is exothermic, the solution is quite hot and it should be allowed to cool
down after mixing. This waiting period is usually 24 hours.

3. After the 24-hour period, the alkaline solution is mixed with fly ash. The quantity of fly
ash is calculated from the water-to-solids ratio given in the first step. The quantities of the
ingredients required to prepare the AAB for jute treatment can be estimated from Table 3.2.
The quantities presented in Table 2 are based on the treatment of unit square metre of jute
geotextile for different ratio of water to solids. It is advisable to consider an additional 10% of
each ingredient to account for wastage during the application stage.

Table 3.2: Amount of AAB required treating per sq. metre jute geotextile

Water/Solid ratio AAB Applied Fly Ash NaOH Sodium Silicate Water
kg/sq. m kg/sq. m kg/sq. m kg/sq. m kg/sq. m
0.35 3.44 2.18 0.058 0.706 0.498
0.40 3.07 1.87 0.049 0.606 0.537
0.45 2.75 1.62 0.043 0.526 0.5611

4. The properties of AAB are similar to that of hydraulic cement. Hence, the hardening process
will start as the alkali solution is added to fly ash. It is advisable to apply the mix while it is
still workable in order to avoid its wastage due to hardening. The mix can either be applied
with paint brushes, or can be sprinkled over the jute geotextile, or jute geotextile can be soaked
such that the mix covers the complete surface area of jute geotextile.

5. The jute geotextile covered with AAB is kept at 40◦ C in a humidity controlled environment
for 24 hours. The treated jute geotextile is then kept at ambient temperature for 7 days to make
it suitable for practical application in ground improvement. This 7-day period enables the AAB
Chapter 3. Materials and Treatment solution 14

Figure 3.1: JGT before and after treatment

to harden and gain sufficient strength to withstand externally applied loads. The treated jute
geotextile is then introduced into the soil in horizontal layers. A series of laboratory experiments
are then conducted to compare the properties of soil, before and after the addition of the treated
jute geotextile.

6. The Fig.3.1 displays the jute geotextile before and after treatment.
Chapter 4

Chemical Characterization of
Treated Jute

Jute is composed of cellulose (58-63%), hemicellulose (20-24%), lignin (12-15%) and traces of oil,
waxes and other impurities(Wang et al., 2009). Due to its composition, jute belongs to a group
of fibres known as lignocellulosic fibres. The following chapter explores the change in chemical
composition, bonds, and mineralogy of jute following the AAB treatment.

4.1 X-Ray Diffraction

The X-Ray diffraction was performed for qualitative identification of chemical structure along
with determining the degree of crystalline nature of untreated and treated geotextile. X-ray
Diffraction analyses are performed using a RIGAKU Ultime IV diffractometer to identify the
minerals present in the untreated and treated jute geotextile. The powdered samples are
examined through CuKα rays generated at 40 mA and 40 kV. The operating 2θ range is from
0o to 100o with a step of 0.02o 2θ and integrated at the rate of 2 seconds per step.

Sharp peaks represented well to perfect crystalline composition while flatter or humped peaks
indicated poor crystallinity. Purely flat portions indicated the presence of amorphous components.
The Fig.4.1 presents the X-Ray diffraction patterns of untreated as well as treated jute samples.
The peaks at 15.7o , 22.3o and 34.5o (denoted by ‘J’) in XRD pattern were exhibited by the
raw jute which was consistent with the study done by Wang et al.(2009) The XRD pattern
of treated jute indicated the presence of quartz (SiO2 ), analcime (N aAlSi2 O6 .H2 O), mullite
(Al6 Si2 O13 ), and hydroxy sodalite (N a6 (Si6 Al6 O24 ).8H2 O) which were recognised by their
characteristic peaks. The existing study by Kar (2013). had shown that these minerals are
peculiar to the hardened AAB paste. Moreover, the peaks characteristic to jute were present

15
Chapter 4. Chemical Characterization of Treated Jute 16

in all the diffractogram of all treated jute samples hence indicating that jute did not lose its
identity during treatment. These concluded that following the treatment hardened paste of AAB
has formed a layer over the jute fibres. This statement was further supported by SEM results.
On a separate note, with increasing water to cement ratio in treatment solution the amorphous
content was observed to be more dominant. This happened due to the production of the greater
amount of sodium aluminosilicate hydrate matrix as a result of greater water content.

4.2 Fourier Transform Infrared Spectroscopy

The FTIR spectra of untreated and treated jute samples were performed to study the change in
the formation of chemical bonds following the treatment by studying the transmittance spectra.
Fourier transform infrared (FTIR) spectroscopy of the untreated and treated fibres are collected
using a JASCO FTIR 4200 setup. The geotextile samples are shredded to powder-like fineness.
The powder is then dried in an oven at 105 ◦ C for 1 h to remove any residual moisture and then
mixed with dried KBr powder to prepare pellets. The spectral range is specified as 4000-400
cm−1 for all the samples. The FTIR spectra for untreated and treated jute samples are performed
to study the change in the formation of chemical bonds following the treatment by analyzing
the transmittance spectra. The Table 4.1 presents the transmittance peaks corresponding to
different bonds.
Table 4.1: Bonds corresponding to characteristic peaks

Bonds Peaks(cm−1 )
O-H stretching 3467-3434
C-H stretching 2988,2921,1381
C-O-C stretching 1035,900
C-O bending 1730
CH2 bending 1428
C=O 1700,1644
C=C in plane 1542
Si-O-Si 1037.5-1024
Al-OH 778-774
Si-O bending 489-469
Si-OH 849

The peaks in Fig.4.2 characterized the bonds present in the untreated jute as well as AAB
treated jute. The transmittance peaks at 3434cm−1 , 3463 cm−1 , 3460 cm−1 , and 3467 cm−1
in untreated ,0.35 w/s AAB treated, 0.40 w/s AAB treated, and 0.45 w/s AAB treated jute,
respectively, represented the O-H stretching absorption. In FTIR spectra of untreated jute, peaks
at 2988 cm−1 , 2921 cm−1 , and 1381 cm−1 represented C-H stretching absorption. The C-O-C
Chapter 4. Chemical Characterization of Treated Jute 17

Figure 4.1: X-Ray Diffractogram for untreated and treated jute sample (J= Jute, S= Hydroxy
sodalite, M= Mullite, Q= Quartz, An= Analcine
Chapter 4. Chemical Characterization of Treated Jute 18

stretching absorptions were observed at 1035 cm−1 and 900 cm−1 . These absorption peaks were
consistent with the FTIR spectra of typical cellulose (Wang et al. 2009). The peaks at 1730
cm−1 and 1428 cm−1 represented the C-O bending and symmetric –CH2 bending vibration
(Abderrahim et al., 2015). The presence of hemicellulose and lignin was also exhibited by the
spectrum. The bands at 1700 cm−1 and 1644 cm−1 correspond to the acetyl groups and C=O
bonds, characteristic of the hemicellulose (Abdulkhani et al., 2013). A lignin peak is located at
the 1542 cm−1 band due to C=C in-plane aromatic vibrations (Neto et al., 2013).

The disappearance of the transmittance peak from 2988 cm−1 to 1700 cm−1 and 1490 cm−1
to 1214 cm−1 indicated that the cellulose backbone and the hemicellulose could have partially
been removed due to the AAB treatment. The peak at 1024 cm−1 , 1034 cm−1 , and 1037.5 cm−1
in 0.35 w/s AAB treated, 0.40 w/s AAB treated and 0.45 w/s AAB treated jute, respectively,
corresponded to the major siloxane unit (Si-O-Si) of the AAB system. Increase in w/s ratio in
the AAB mixes led to the chemical shift observed in case of the Si-O-Si peaks. The respective
peaks for treated jute samples at 774 cm−1 , 777cm−1 , and 778 cm−1 corresponded to Al-OH
stretching vibration. The peaks at 520 cm−1 in case of 0.35 w/s AAB treated jute represented
Si-O-Al bending vibration. The peaks in treated jute samples at 483–469 cm−1 represented
Si-O in-plane bending vibration. The peak at the 849 cm−1 in 0.45 w/s AAB treated jute
represented the bending vibration of the Si-OH bond. These peaks represent the aluminosilicate
components of the AAB system (Kar, 2013). However, almost same transmittance peaks were
observed in the FTIR spectrum of the AAB treated jute samples as that of untreated jute sample.
This indicated that the structure of cellulose had not been completely damaged following the
treatment. Moreover, transmittance peaks at 1644 cm−1 ,1647 cm−1 and 1644 cm−1 in AAB
treated samples corresponded to the acetyl groups and C=O bonds of hemicellulose. This shows
that the jute did not lose its identity completely after treatment.

4.3 Surface Morphology

The SEM-EDS study is conducted using the JSM-7600F, a thermal Field Emission Scanning
Electron Microscope (FE-SEM), provided by JEOL Ltd. The jute geotextiles used in the study
are mostly composed of cellulose, hemicellulose. In case of treated jute geotextiles, there is the
additional presence of fly ash. Hence, the possibility of the presence of elements having atomic
numbers higher than that of iron is rare in this case. However, if the raw materials of treatment
solution or jute are collected from an unknown source, a thorough investigation has to be carried
out. In that case, 20kV might be used as excitation voltage. However, the high excitation voltage
would cause an electron cloud, hence the obtained image would be blurred. The commonly used
range for alkali activated binder is 5kV–15 kV (Kar, 2013). Hence, the excitation voltage for
this study is kept as 10kV. The images of different magnification are captured, at three different
Chapter 4. Chemical Characterization of Treated Jute 19

Figure 4.2: FTIR spectral pattern of (a) Untreated jute, (b) 0.35 w/s AAB treated jute, (c)
0.40 w/s AAB treated jute, and (d) 0.45 w/s AAB treated jute.

locations for each sample. At each location, three different regions are chosen at random and
further in each of these regions, five different points are studied. The EDS spectra are provided
by INCA software system. For the adequate working of EDS analyzer, the probe current and
working distance are maintained at 65.4 - 67.0 µA and 8-15 mm, respectively. Prior to analysis,
the geotextiles fibres are dried at 105 ◦ C for 24 hours to remove all the internal moisture.To
make the fibres electrically conductive, a 15nm layer of platinum is coated on the jute fibres in
argon gas atmosphere.

The Fig.4.3, Fig.4.4, Fig.4.5 and Fig.4.6 , respectively, shows the images of untreated jute, 0.35
w/s ratio treated jute, 0.40 w/s ratio treated jute, and 0.45 w/s ratio treated jute, each at the
X2000 as well as X5000 magnification along with EDS analysis. Several grooves are visible in
the SEM micrographs along the longitudinal section of the jute fibres in each case. The surface
of the raw jute fibre is covered by impurities like hemicellulose, lignin, pectin (Wang et al.,
2009). The 0.35 w/s AAB-treated jute fibre showed unevenly distributed cavities shows that the
fibre surface gets partially covered with a matrix of sodium aluminosilicate hydrate (N-A-S-H),
characterized by vitreous networks. Additionally, there is some unreacted residue of fly ash,
Chapter 4. Chemical Characterization of Treated Jute 20

distinguished by the spherical particles (Fig. 4.3). As the w/s in the treatment mix increase, the
extent of coating on the jute fibre increases. Due to greater w/s ratio, there is a greater extent
of reaction between fly ash and the alkali-activating solution. Hence, the figures (Fig. 4.4-4.6)
show an increase in the proportion of the N-A-S-H matrix and a consequent reduction in the
unreacted fly ash residue.

EDS shows the predominance of C and O in the untreated jute (Fig. 4.3-4.6), as this technique
is unable to detect any element below Be in the periodic table. Hence, the presence of H is
not seen. There are traces of Na and Si. Al and Fe are completely absent, as expected. In
case of the treated jute specimens, the marked presence of Al and Si are noticed due to their
abundance in fly ash. Increase in Na content is also noticed due to its substantial presence in
the alkali-activating solution. The decrease in C-content in treated jute shows the dominance of
AAB characteristics over jute properties to some extent. Since EDS scans the whole volume
of the specimen and not just the surface, it cannot be inferred that AAB just adheres to the
surface of the jute fibres and leads merely to changes in surface characteristics. The results
provide indication towards the possible formation of a jute-AAB composite. These findings are
corroborated by the results from tensile strength experiment.

4.4 Thermogravimetric Analysis

Differential thermogravimetric analysis (TGA) is carried out through a SHIMADZU/DTG-60


setup to evaluate the thermal stability of untreated and treated jute. The geotextile samples
weighing between 4 and 10 mg are kept on a platinum pan in a nitrogen-rich environment. The
samples are heated gradually from 30 ◦ C to 950 ◦ C, which is in agreement with the decomposing
range of jute (Yang et al., 2007). The rate of temperature increment is kept as 10 ◦ C/min. to
assure uniform heating.

The thermogravimetric curve in Fig.4.7 provided the information regarding the thermal stability
of untreated and treated jute. As the temperature increased up to 950 ◦ C, various components in
sample started to decompose providing the change in weights hence yielding the identity of the
present components. The TG curve of untreated jute sample represented the initial loss in weight
at 100 ◦ C due to vaporization of the evaporable water. The thermal degradation of biomass
in untreated jute took place in three phases- thermal decomposition of lignin, hemicellulose
and cellulose. The weight loss in untreated jute from 270 ◦ C to 310 ◦ C was a consequence of
the hemicellulose decomposition. Following the hemicellulose decomposition, major weight loss
occurred within the temperature range of 323-392 ◦ C due to cellulose decomposition and the
process was further extended up to 525 ◦ C. Lignin decomposition started at the temperature
range of 155-169 ◦ C, however, lignin is constituted of various aromatic branches that provide high
thermal stability to lignin compared to cellulose and hemicellulose. Therefore, the decomposition
Chapter 4. Chemical Characterization of Treated Jute 21

Figure 4.3: SEM images of untreated jute at X2000 and X5000 magnification along with EDS
analysis
Chapter 4. Chemical Characterization of Treated Jute 22

Figure 4.4: SEM images of 0.35 AAB treated jute at X2000 and X5000 magnification along
with EDS analysis
Chapter 4. Chemical Characterization of Treated Jute 23

Figure 4.5: SEM images of 0.40 AAB treated jute at X2000 and X5000 magnification along
with EDS analysis
Chapter 4. Chemical Characterization of Treated Jute 24

Figure 4.6: SEM images of 0.45 AAB treated jute at X2000 and X5000 magnification along
with EDS analysis
Chapter 4. Chemical Characterization of Treated Jute 25

of lignin took place in the wide temperature range of 150-550 ◦ C. The similar behaviour was
observed by Yang et al.(2007) and they reported that the decomposition of cellulose, hemicellulose
and lignin occurred in the range of 315-400 ◦ C, 220-315 ◦ C, and 100-900 ◦ C, respectively. An
alternative study by Nunn et al.(1985) showed the range of cellulose and lignin decomposition to
be 200-400 ◦ C and 150-750 ◦ C, respectively.

Figure 4.7: Thermogravimetric curves of untreated and AAB treated jute

The jute treated with various AAB solutions showed the combined characteristic of both jute
and AAB. A study by Bakri et al.(2012) had shown that the AAB lost evaporable water up to
250 ◦ C followed by a constant decrease in weight up to 700 ◦ C. Moreover, the change in weight
was observed to be negligible after 700 ◦ C.The consistent results were observed for the treated
jute samples. The TG curves of treated samples show the rapid change in weight up to 250 ◦ C
due to loss of evaporable water. It is followed by the loss of lignin, hemicellulose and cellulose of
jute along with a constant reduction in weight of AAB. Above 700 ◦ C, little change in weight
was observed for each of the treated jute specimens. This showed that the fabric lost it complete
integrity after 700 ◦ C and only was left behind.
Chapter 5

Mechanical Properties of treated


geotextile

Figure 5.1: Tensile strength of raw jute and treated jute geotextile

The tensile strength of narrow strip of geotextile was determined by “MCS” UTM. The width
and the gauge length of a tested sample were 50mm and 200mm, respectively. The grab length
was kept as 25mm. The tensile loading was maintained at a deformation rate of 100 mm/min
following the ISO 13934-1:1999 standards to determine narrow strip tensile strength. Ten

26
Chapter 5. Mechanical Properties of treated geotextile 27

specimens of each type of geotextile were tested in the machine direction and the average of
tensile strength was reported for each sample.

The tensile strength of jute was increased remarkably following the treatment with AAB solution.
The tensile strength of untreated and treated jute samples are presented in Fig. 5.1. The tensile
strength of treated jute sample was measured after 28 days following the application of solution
as the AAB gained 95% of its compressive strength after 28 days ( Kar, 2013). It was also
assured that the relative humidity for all the samples was kept constant while curing. The tensile
strength of jute was increased by 35.7%, 35.8%, and 34% after treating it with AAB solution
of 0.35, 0.40, and 0.45 w/s ratio, respectively. A conceivable explanation is that, following the
treatment, the combined stiffness of jute-AAB composite is higher than that values of jute and
AAB. It was also observed that the modulus of elasticity of jute and hardened AAB lies in
the range of 20.4-24.2 GPa (Mwaikambo, 2009) and 18.7-21.4 GPa(Kar, 2013), respectively.
Consequently, the treatment of jute with AAB lead to the increased stiffness that further led to
improvement in the load-bearing capacity of jute.
Chapter 6

Sand reinforced with treated jute

6.1 Vertical Permeability

The vertical permeability of reinforced sand was determined by the falling head permeability test
(ASTM D2434-68 (2006) and IS 2720 – 1986 Part 17). The steel cylindrical mould of diameter
100 mm and depth 115 mm was used to enclose the soil. The reinforcement layer was placed
at the depth of 50 mm such that it perfectly covers the horizontal surface area. The porous
stone was placed at the top of the cylindrical mould for uniform distribution of water over the
horizontal cross-section of the mould. A layer of filter paper was provided at the bottom of the
mould to avoid the erosion of soil hence consequently avoiding blockage of the discharge pipe.
The known weight of sand was filled layer by layer such that the relative density of 40% was
maintained. The tests were performed 20 times for 3 samples of each type of reinforcement to
abate the error and the mean values were considered as the final vertical permeability of the
reinforced soil.

It was observed (Fig.6.1) that there was a slight decrease in the vertical permeability of
sand reinforced with treated geotextile, however, the change was insignificant as the range
of permeability for all reinforcement lies in the same order of 10-2 cm/s. The results inferred
that the inclusion of a layer of treated jute geotextile did not affect the high permeability of sand
adversely. The permeability results showed the presence of voids in jute geotextile, following the
treatment, that allows the seepage of water.

6.2 Interface properties of sand and reinforcement

The “Heico” automatic large-size direct shear apparatus was used to study the interface properties
of soil and reinforcement. The experiment was carried out according to the ASTM D5321. The

28
Chapter 6. Sand reinforced with treated jute 29

Figure 6.1: Permeability of reinforced sand

upper shear box of size 300mm length, 300mm width and 150mm depth was filled with sand
and the geotextile was fixed at the top of the lower shear box. The set of gears, powered by an
electric motor, controlled the motion of the lower shear box. The thick rigid plate was placed at
the top of the upper shear box to uniformly distribute the normal load. The normal load was
applied by a hydraulic jack that transferred it to the rigid plate through a reaction frame. The
horizontal and vertical loading capacities of apparatus were 100 kN and 100 kN, respectively.
The horizontal and vertical displacements were measured by LVDTs of 100mm capacities. The
shear rate was kept as the 1.2617 mm/min.

The interface friction angles between sand and geotextile reinforcement are provided in the
Fig.6.2. It was observed that the interface friction between untreated jute and sand was 39.1%
higher than the internal angle of friction of sand. However, the treatment of jute significantly
abated the interface friction between sand and geotextile. The interface friction values between
jute and sand were dropped by 32.5%, 38.3%, and 45.7%, respectively, following the treatment
with AAB solution of 0.35, 0.40, and 0.45 w/s ratios. On the contrary, the cohesion was developed
between geotextile and sand following the treatment. It can be observed from Fig.6.3, that the
cohesion between jute and reinforcement is commensurate to the water to solid ratio of the
treatment solution.

A separate study is also conducted to study the effect of strain rate on the interface friction angle
of jute. In this experiment, the relative density of sand is maintained at 50%. The interface
friction angles are observed at three different strain rates, 1.01 mm/min, 1.58 mm/min, and
Chapter 6. Sand reinforced with treated jute 30

Figure 6.2: Friction angle between sand and reinforcement

Figure 6.3: Cohesion between sand and reinforcement


Chapter 6. Sand reinforced with treated jute 31

1.778 mm/min. The 45 cycles of large shear box test are conducted for this study. Fig. 6.4 shows
the interface friction angle corresponding to different shear strain rates. It is observed that the
friction angle decreases slightly with an increase in strain rate. The reduction in friction angle is
generally observed to be linear. This trend of reduction in friction angle between geosynthetics
and sand is also observed by Mamo and Dey (2010). Moreover, the friction angle is highest for
untreated jute, followed by 0.35 AAB treated jute, 0.40 AAB treated jute, and 0.45 AAB treated
jute. This trend is similar to the last sets of results.

Figure 6.4: Interface Fricion angle vs Shear Strain Rate

6.3 Model Plate Load Test

The relative study of the bearing capacity of sandy soil reinforced with treated and untreated
geotextile was done in the modelled plate load setup. The tests were conducted in a steel tank
of size 1.2m length, 0.91 m width and 0.91 m depth. The loading setup was fabricated of steel
columns and channels that supported the hydraulic loading cell of 150 kN capacity. The loading
cell transfers the load to the soil through the square steel plate of side 300 mm and 25 mm depth.
The plate was kept at the centre and it was ensured that it remained horizontal for efficient
transfer of load. The settlement of the plate was measured by the four Linear variable differential
transformers (LVDT) placed at the radial distance of 210 mm from the centre of the load cell.
The capacities of LVDTs were 50mm. The Fig.6.5 shows the real setup of the plate load.
Chapter 6. Sand reinforced with treated jute 32

Figure 6.5: Experimental Setup of model plate load

The sand was filled layer by layer and uniformly compacted to a relative density of 40%. The
compaction of sand was carried out by 10 kg hammer falling from the height of 20 cm. The layer
of geotextile was placed at the depth of 100 mm, 150mm, 200mm, and 250mm. The width of the
roll of geotextile was equal to the width of the tank, therefore, a single piece of geotextile was
cut from the roll and used to cover the whole surface area. As per IS: 1888 (1982), the strain
was maintained at a constant rate of 0.02 mm/min. Loading was continued until the settlement
reached a value of 50mm.

The load-settlement curves of the only sand and sand reinforced with untreated as well as treated
jute geotextile are shown in Fig. 6.6, Fig. 6.7, Fig.6.8, and Fig. 6.9. The inclusion of untreated
jute geotextile in sand improved the loading capacity slightly, however its presence increased
the displacement corresponding to ultimate load. On a separate note, the ultimate bearing load
capacity, as well as displacement corresponding to the ultimate load, were increased as sand was
reinforced with treated jute, nevertheless, the degree of the influence on the latter properties
Chapter 6. Sand reinforced with treated jute 33

Figure 6.6: Load vs Displacement graph at the depth of 10 cm

Figure 6.7: Load vs Displacement graph at the depth of 15 cm


Chapter 6. Sand reinforced with treated jute 34

Figure 6.8: Load vs Displacement graph at the depth of 20 cm

Figure 6.9: Load vs Displacement graph at the depth of 25 cm


Chapter 6. Sand reinforced with treated jute 35

significantly varied with the water to solid ratio of the treatment mix. The maximum ultimate
load was observed for the sand reinforced with jute treated with a solution of 0.35 w/s ratio and
was followed by the jute treated with a solution of 0.40 and 0.45 w/s ratio, respectively. However,
the trend shown by the displacement corresponding to ultimate load was quite contrasting
as its value decreased with increasing water to solid ratio. At the depth of 100mm, all the
reinforcements samples were ruptured. It was also observed that the layer of reinforcement at
the depth of 25cm don’t have significant effect on the bearing capacity.
Chapter 7

Numerical Study

A numerical model is a mathematical simulation to a real physical problem. The results obtained
from the plate load test are simulated in commercially available software PLAXIS 2D 2016.
A sand chamber of the same dimensions is modelled for the unreinforced sand and the sand
reinforced with untreated and different treated jute geotextile. As the experimental plate load
results gave optimum results for the depth of 10 cm, the numerical model for that is studied in
the following section.

To keep the similar boundary conditions as the model plate load, the bottom edge is kept fixed
in both horizontal and vertical direction and the vertical side edge is kept fixed in horizontal
direction. The vertical edge of the box is kept at the adequate distance from the edge of the
footing so that the results are not affected by the side edges. The geostatic stress is kept as the
initial condition for the numerical model. The stage construction is used in the modelling, where
the displacement is applied step by step at the rate of 1mm/sec on the footing. The vertical
reaction load corresponding to each displacement is calculated.

7.1 Footing

In the present study, as the model is axisymmetric, the circular plate is used for the application
of load. The base area of the circular plate is kept same as that of the square plate used in the
experimental model plate load. Hence, the equivalent circular plate of diameter 225 mm and
depth 25 mm has been used in the numerical model (Fig. 7.1-7.2). The plate is considered as
rigid and rough.

36
Chapter 8. Numerical Study 37

Figure 7.1: Meshing with (a) Soil (b) Soil reinforced with geotextile

7.2 Soil and Geotextile

The elastic-perfectly plastic Mohr-Coulomb model is used to determine the behaviour of reinforced
sand under the load. The model used in this study is consistent with existing literature on
foundation design (Potts and Zdravkovic, 2001). 12 point integrated, 15 noded triangular
elements are used to model the soil. The modulus of elasticity is taken as 23,000 kPa which lies
within the typical range of the modulus of elasticity of sand (Bowles 1988). The Poisson’s ratio
of sand was taken as 0.3 for all the simulations. The axial stiffness of geotextile is calculated from
the stress-strain graph, obtained from the tensile test. The values of axial stiffness of untreated,
0.35 w/s AAB treated, 0.40 w/s AAB treated, and 0.45 w/s AAB treated jute sample are 220
kN/m, 373.3 kN/m, 280 kN/m, and 240 kN/m respectively.

7.3 Mesh

The meshing or discretisation in finite element model is a process of breaking the numerical
model into smaller discrete pieces. In the present study, very fine global coarseness along with
cluster and global refinement are used. The former refinement is used to intensify the number of
elements in particular cluster and the latter refinement is used to globally increase the number of
elements. To intensify the elements in the boundaries, line refinement is done for the rigid plate.
Additionally, the dense and smaller meshing is done below the plate as the stress-strain behaviour
under the plate is of primary importance. The meshing details of the , both undeformed and
deformed, unreinforced and reinforced sand are provided in Fig. 7.1 and Fig. 7.2, respectively.
Chapter 8. Numerical Study 38

Figure 7.2: Deformed mesh for (a) Soil, (b) Soil reinforced with AAB treated jute geotextile.

Figure 7.3: Comparison of load-settlement curves from numerical and experimental study.
Chapter 8. Numerical Study 39

7.4 Results

The comparison of load-settlement curves from experimental and numerical methods is presented
in Fig. 7.3. The results obtained from the numerical study show a good resemblance with the
results obtained from the experimental study. It is observed that the load-settlement curves
are roughly same, irrespective of geotextile, that is also observed from the experimental study.
Hence, it may be concluded that the numerical study is in good agreement with the experimental
study and so verifies the experimental results. The further study can be done on the basis of the
present model.
Chapter 8

Durability Study

The jute treated with 0.40 AAB solution and untreated jute were kept in the sand with high
water content of 12% for the period of 1 month and 3 months. The sand was stored in the
closed container to maintain the relative humidity and was changed after each week to ensure
the uniformity of saturation. The Fig. 8.1 shows both jute samples after the time of exposure.

Figure 8.1: Treated and untreated jute exposed to unsterile soil (1 month)

The Fig 8.1 shows that the untreated jute was severely damaged and started to degrade. However,
the treated jute had complete integrity and there was no visible damage. The tensile strength
tests are conducted on the samples after exposing them to the unsterile soil. The results obtained
are remarkable. After 3 months, jute is completely disintegrated, however, treated jute samples
retained 44-58 % of its strength. Fig. 8.2 shows the tensile strength of treated and untreated jute

40
Chapter 8. Durability Study 41

after exposure corresponding to the time of exposure. The possible reason for high durability of
treated jute samples is a layer of AAB on jute geotextile that is highly resistant to biological
and chemical degradation.

Figure 8.2: Tensile strength of geotextile after exposure to unsterile soil


Chapter 9

Conclusion

The natural geotextile is more suitable than synthetic geotextiles as they don’t leave a carbon
footprint and cost less. Moreover, the use of fly ash as treatment ingredient serves dual benefit
of aiding environment by avoiding the disposal of fly ash and decreasing the cost associated to
fly ash disposal. The current techniques of treating jute to improve its strength, durability, and
performance, require convoluted methods and expensive chemicals. Therefore, the present study
devises a new technique of treating jute geotextile with AAB solution of different w/s ratio
(0.35-0.45) in order to improve the strength characteristics of the soil. The important findings of
the present study can be summarized as follows:

• FTIR and TGA show that the characteristic of untreated jute is more or less reserved
following the treatment with AAB coating. With increasing, w/s ratio, the amorphous
character of sodium aluminate silicate hydrate from the alkali activated fly ash becomes
predominant as deduced from XRD results.

• SEM/EDS results show the production of a greater quantity of sodium aluminosilicate


hydrate at higher w/s ratios. These aluminosilicates contained a greater number of pores
which leads to higher permeability. On the contrary, there is the relatively lesser quantity
of this aluminosilicate at lower w/s ratio.

• Consequently, this leads to lesser porosity and hence lower permeability and higher load-
bearing capacity. These findings are later corroborated by permeability test and tensile
strength test.

• The surface roughness of jute fabric decreases to some extent as a consequence of AAB
coating. Hence the shear strength of treated jute decreases slightly. Additionally, with
increasing shear strain rate, a decrease in interface friction angle is observed.

42
Chapter 9. Conclusion 43

• The layer of AAB coating becomes gradually softer with increasing w/s ratio, thus reducing
the flexibility of jute geotextile. This leads to a lower bearing capacity of jute fabric treated
with higher w/s AAB ratio. A w/s ratio of 0.35 is found to be optimum for achieving
maximum load-bearing capacity.

• The numerical analysis shows a similar trend for bearing capacity; however, the peak load
is more clearly defined in the experimental results compared to numerical study.

• The durability of jute improved substantially following the treatment with Alkali activated
Binder.
Chapter 10

References

Abdulkhani, A., Marvast, E.H., Ashori, A., Hamzeh, Y. and Karimi, A.N., 2013. Preparation
of cellulose/polyvinyl alcohol biocomposite films using 1-n-butyl-3-methylimidazolium chloride.
International journal of biological macromolecules, 62, pp.379-386.

Abderrahim, B., Abderrahman, E., Mohamed, A., Fatima, T., Abdesselam, T. and Krim, O.,
2015. Kinetic thermal degradation of cellulose, polybutylene succinate and a green composite:
comparative study. World Journal of Environmental Engineering, 3(4), pp.95-110.

Ahmad, F., Bateni, F. and Azmi, M., 2010. Performance evaluation of silty sand reinforced with
fibres. Geotextiles and Geomembranes, 28(1), pp.93-99.

Al Bakri, A.M., Kamarudin, H., Bnhussain, M., Nizar, I.K., Rafiza, A.R. and Zarina, Y., 2012.
The processing, characterization, and properties of fly ash based geopolymer concrete. Rev. Adv.
Mater. Sci, 30, pp.90-97.

Andersson, M. and Tillman, A.M., 1989. Acetylation of jute: Effects on strength, rot resistance,
and hydrophobicity. Journal of applied polymer science, 37(12), pp.3437-3447.

ASTM D1195 / D1195M-09(2015), Standard Test Method for Repetitive Static Plate Load Tests
of Soils and Flexible Pavement Components, for Use in Evaluation and Design of Airport and
Highway Pavements, ASTM International, West Conshohocken, PA, 2015, www.astm.org.

ASTM D2434-68(2006), Standard Test Method for Permeability of Granular Soils (Constant
Head) (Withdrawn 2015), ASTM International, West Conshohocken, PA, 2006, www.astm.org.

ASTM D5321 / D5321M-17, Standard Test Method for Determining the Shear Strength of Soil-
Geosynthetic and Geosynthetic-Geosynthetic Interfaces by Direct Shear, ASTM International,
West Conshohocken, PA, 2017, www.astm.org.

44
Chapter 10. References 45

Basu, G., Roy, A.N., Bhattacharyya, S.K. and Ghosh, S.K., 2009. Construction of unpaved rural
road using jute–synthetic blended woven geotextile–A case study. Geotextiles and Geomembranes,
27(6), pp.506-512.

Datta, U., 2007. Application of jute geotextiles. Journal of Natural Fibers, 4(3), pp.67-82.

Dey, A., STRAIN RATE AND SCALE EFFECTS ON STRESS-STRAIN BEHAVIOR OF


SAND.

Indian Standard: 1888–1982 Method of load test on soils, Bureau of Indian Standards, New
Delhi

Indian Standard: 2720 Methods of tests for soils (Part 14)—1983, Determination of density
index (relative density) of cohesionless soils, Bureau of Indian Standards, New Delhi

Indian Standard: 2720 Methods of tests for soils (Part 17)—1986, Determination of permeability
of soil, Bureau of Indian Standards, New Delhi

ISO, E., 1999. 13934-1: 1999. Textiles-Tensile properties of fabrics-Part, 1, pp.13934-1

Jual G.P. and Sharda V.N., 2008. Eco-Rehabilitation of Himalayan Hill slopes, affected mine
spoils and landslide using Jute Geotextile, In Proc International Workshop on Jute Geotextile,
pp. 56-57, 2008

Kar, A., 2013. Characterizations of concretes with alkali-activated binder and correlating their
properties from micro-to specimen level. West Virginia University.

Li, J., Tang, C., Wang, D., Pei, X. and Shi, B., 2014. Effect of discrete fibre reinforcement on
soil tensile strength. Journal of Rock Mechanics and Geotechnical Engineering, 6(2), pp.133-137.

Maione, U., Majone-Lehto, B. and Monti, R. eds., 2000. New trends in water and environmental
engineering for safety and life. CRC Press.

Mwaikambo, L.Y., 2009. Tensile properties of alkalised jute fibres.

Nam, S. and Netravali, A.N., 2006. Green composites. I. Physical properties of ramie fibers for
environment-friendly green composites. Fibers and Polymers, 7(4), pp.372-379.

Neto, A.R.S., Araujo, M.A., Souza, F.V., Mattoso, L.H. and Marconcini, J.M., 2013. Characteri-
zation and comparative evaluation of thermal, structural, chemical, mechanical and morphological
properties of six pineapple leaf fiber varieties for use in composites. Industrial Crops and Products,
43, pp.529-537.

Nunn, T.R., Howard, J.B., Longwell, J.P. and Peters, W.A., 1985. Product compositions and
kinetics in the rapid pyrolysis of sweet gum hardwood. Industrial & Engineering Chemistry
Process Design and Development, 24(3), pp.836-844.
Chapter 10. References 46

Prodhan Z. H., 2008. Application of Jute Geotextile on Rural Roads of Bangladesh for slope
Protection. In Proc International Workshop on Jute Geotextile, pp. 49-52.

Rao, A.S., 2003, August. Jute geotextile application in kakinada port area. In Proc. Nat.
Seminar on JuteGeotextile & Innovative jute products.

Rawal, A. and Sayeed, M.M.A., 2014. Tailoring the structure and properties of jute blended
nonwoven geotextiles via alkali treatment of jute fibers. Materials & Design, 53, pp.701-705.

Rujikiatkamjorn, C., Indraratna, B.,and Chu, J. 2015. Ground Improvement- Case Studies.
Elsevier

Saha, P., Roy, D., Manna, S., Adhikari, B., Sen, R. and Roy, S., 2012. Durability of transesterified
jute geotextiles. Geotextiles and Geomembranes, 35, pp.69-75.

Sanyal, T. and Chakraborty, K., 1994. Application of a bitumen-coated jute geotextile in bank-
protection works in the Hooghly estuary. Geotextiles and Geomembranes, 13(2), pp.127-132.

Sanyal, T., 2017. Jute Geotextiles and Their Applications in Civil Engineering. Springer.

Wambua, P., Ivens, J. and Verpoest, I., 2003. Natural fibres: can they replace glass in fibre
reinforced plastics?. composites science and technology, 63(9), pp.1259-1264.

Wang, W.M., Cai, Z.S., Yu, J.Y. and Xia, Z.P., 2009. Changes in composition, structure, and
properties of jute fibers after chemical treatments. Fibers and Polymers, 10(6), pp.776-780.

Yang, H., Yan, R., Chen, H., Lee, D.H. and Zheng, C., 2007. Characteristics of hemicellulose,
cellulose and lignin pyrolysis. Fuel, 86(12), pp.1781-1788.
Appendix A

MATLAB Code for Curve Fitting

The following code is used to fit a smooth polynomial curve, that delineates the plate load
results.

s1 = polyf it(SD , SL , 8);

sf it = polyval(s1, SD );

plot(SD , sf it,0 +0 );

holdon

u1 = polyf it(UD , UL , 8);

uf it = polyval(u1, UD );

plot(UD , uf it,0 −−0 );

holdon

j351 = polyf it(J35D , J35L , 8);

j351 f it = polyval(j351 , J35D );

plot(J35D , j351 f it,0 .0 );

holdon

j401 = polyf it(J40D , J40L , 8);

j401 f it = polyval(j401 , J40D );

plot(J40D , j401 f it,0 :0 );

holdon
47
Appendix A. MATLAB Code for Curve Fitting 48

j451 = polyf it(J45D , J45L , 8);

j451 f it = polyval(j451 , J45D );

plot(J45D , j451 f it,0 −.0 );

xlabel(0 Displacement(mm)0 ); ylabel(0 Load(kN )0 ); xlim([050]); ylim([0120]); legend(0 Sand0 ,0 Sand+


U ntreatedJute0 ,0 Sand+0.35w/sAABT reatedJute0 ,0 Sand+0.40w/sAABT reatedJute0 ,0 Sand+
0.45w/sAABT reatedJute0 );

x0 = 10;

y0 = 10;

width = 600;

height = 250;

set(gcf,0 units0 ,0 points0 ,0 position0 , [x0, y0, width, height])

Вам также может понравиться