You are on page 1of 31

Journal Pre-proof

Comparative study of pineapple leaf microfiber and aramid fiber reinforced natural
rubbers using dynamic mechanical analysis

Budsaraporn Surajarusarn, Samar Hajjar-Garreau, Gautier Schrodj, Karine Mougin,


Taweechai Amornsakchai

PII: S0142-9418(19)31896-3
DOI: https://doi.org/10.1016/j.polymertesting.2019.106289
Reference: POTE 106289

To appear in: Polymer Testing

Received Date: 16 October 2019


Revised Date: 3 December 2019
Accepted Date: 10 December 2019

Please cite this article as: B. Surajarusarn, S. Hajjar-Garreau, G. Schrodj, K. Mougin, T.


Amornsakchai, Comparative study of pineapple leaf microfiber and aramid fiber reinforced natural
rubbers using dynamic mechanical analysis, Polymer Testing (2020), doi: https://doi.org/10.1016/
j.polymertesting.2019.106289.

This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition
of a cover page and metadata, and formatting for readability, but it is not yet the definitive version of
record. This version will undergo additional copyediting, typesetting and review before it is published
in its final form, but we are providing this version to give early visibility of the article. Please note that,
during the production process, errors may be discovered which could affect the content, and all legal
disclaimers that apply to the journal pertain.

© 2019 Published by Elsevier Ltd.


Comparative Study of Pineapple Leaf Microfiber and Aramid Fiber Reinforced

Natural Rubbers using Dynamic Mechanical Analysis

Budsaraporn Surajarusarna,b, Samar Hajjar-Garreaub, Gautier Schrodjb,

Karine Mouginb and Taweechai Amornsakchaia,c*,


a
Polymer Science and Technology Program, Department of Chemistry and Center of

Excellence for Innovation in Chemistry, Faculty of Science, Mahidol University,

Phuttamonthon 4 Road, Salaya, Phuttamonthon District,

Nakhon Pathom 73170, Thailand


b
Institut de Science des Matériaux de Mulhouse, IS2M-CNRS-UHA, 15, Rue Jean Starcky,

B.P.2488-68057 Mulhouse Cedex, France


c
Center of Sustainable Energy and Green Materials, Faculty of Science, Mahidol University,

Phuttamonthon 4 Road, Salaya, Phuttamonthon District, Nakhon Pathom 73170, Thailand

* Corresponding author

Email: taweechai.amo@mahidol.edu

1
Abstract

Natural fiber is often considered inadequate for high performance reinforcement of

polymer matrix composites. However, some natural fibers have relatively high mechanical

properties with modulus close to that of high-performance synthetic fibers. Since the

reinforcing efficiency of a short fiber is determined not only by the fiber modulus, but also by

other physical properties such as the length to diameter ratio. Here it is shown, for the first

time, that pineapple leaf fiber, whose modulus is somewhat lower than that of aramid fiber,

can be used to reinforce natural rubber more effectively than aramid fiber. The situation was

achieved by breaking down the fiber bundles into the constituent microfibers to gain very

high aspect ratio. Comparisons were made at fiber contents of 2, 5 and 10 parts (by weight)

per hundred of rubber (phr) using dynamic mechanical analysis over a range of temperature.

The results reveals that at temperature below the glass transition of the matrix rubber and low

fiber contents of 2 and 5 phrs, aramid fiber displays slightly better reinforcement efficiency.

At high temperatures of 25 and 60 °C and high fiber content of 10 phr, pineapple leaf

microfiber clearly displays higher reinforcement efficiency than does aramid fiber. Surface

modification of the fiber by silane treatment provides a slight improvement in reinforcing

efficiency.

Keywords: fiber reinforced rubber; natural rubber composites; pineapple leaf fiber; adhesion

2
1. Introduction

A great deal of attention has been paid to the use of natural fibers as reinforcement for

polymer matrix composites. Natural fiber reinforced plastics are studied much more than are

rubbers and have reached commercial production, see for example [1]. The major problem

with high extensibility of rubbers is that failure is more sensitive to the size of the filler

incorporated [2,3]. There have been several attempts to use natural fibers to reinforce rubbers

[4 - 13]. We have shown that pineapple leaf fiber (PALF), whose size is relatively small

compared to other natural fibers, can be used to reinforce rubber effectively [14]. However,

the composites fail at rather low elongation at break. Only when PALF is combined with an

appropriate hybrid particulate filler can the elongation at break of the composite be extended

remarkably [15,16]. In addition, the stress-strain curves of PALF- reinforced rubbers can be

manipulated by way of different techniques such as an increase in the rubber mastication time

[17], controlling the distribution of hybrid particulate filler [18], adjusting the degree of

crosslinking [19] or adding polar rubber [20]. Despite the different approaches that have

been tried, all the composites display failure with fiber pull-out. So there is still some room

for improvement. Rubber reinforced with aramid fiber has been studied as its properties can

be used as a benchmark [21]. Such composites show excellent mechanical properties without

any sign of fiber pull-out. For reinforcement with only aramid fiber, the rubber composites

break at relatively low elongations. It has been confirmed that only with a combination of

hybrid particulate filler do the aramid fiber reinforced rubber composites have high

elongation at break. So, with present knowledge, PALF cannot compete with aramid fiber as

reinforcing agent in rubber composites. This is especially the case when comparison is made

over large strains or deformations when good interfacial adhesion is required. In order to

evaluate the true potential of PALF itself, it is important to compare the composite properties

at very low strains when interfacial adhesion and stress transfer are still effective.

3
According to the short fiber reinforced composite model, the composite modulus is

determined by both the fiber’s modulus and its aspect ratio. For PALF, it is possible to

prepare the fiber so as to have a very high aspect ratio and this could, in effect, compensate

for its lower modulus.

The objective of this paper is to compare the reinforcing efficiency of aramid and

PALF in rubber matrix composites at very low strain over a temperature range using a

dynamic mechanical analysis technique. PALF is normally in a form of bundles of small

elementary fibers that can be easily separated using sodium hydroxide solution [22]. So far

PALF reinforced rubbers show lower moduli than do aramid fiber reinforced rubbers (cf. [16]

and [21]). To get the best effect of PALF, the fiber bundles need to be separated into their

elementary micron size form while keeping their length the same. This will be called

pineapple leaf microfiber (PALMF). The effect of was compared at fiber loadings of 2, 5 and

10 In addition, surface treatment of the PALMF with silane was carried out in order to

improve the interfacial adhesion.

2. Experimental

2.1 Materials

Rubber and chemicals

Natural rubber (NR) was STR5L grade and was purchased from M.B.J. Enterprise

Co., Ltd. (Bangkok, Thailand). The curing additives, i.e., zinc oxide (ZnO), stearic acid,

sulfur and CBS (N-Cyclohexyl-2-benzothaiazolesulfenamide) were commercial grade. The

aramid fiber was Kevlar® (DuPont, USA), supplied by Polymer Innovation Co., Ltd.

(Bangkok, Thailand). It was supplied in the form of a masterbatch in rubber with a fiber

content of 40 wt% Bis-[γ-(triethoxysilyl)-propyl]-tetrasulfide (silane-69) was purchased from

4
Sigma-Aldrich. Sodium hydroxide (NaOH) was purchased from Alfa Aesar. Toluene and

ethanol were purchased from Carlo Erba.

Pineapple leaf microfiber (PALMF)

Pineapple leaf fiber (PALF) was obtained from Bang Yang District, Phitsanulok

Province, Thailand. The process for separating pineapple leaf fiber followed the method

proposed by Kengkhetkit and Amornsakchai [23]. PALF was treated with 10% w/v NaOH

solution for 30 min. Then, it was stirred using a homogenizer for 30 min. The fiber was

washed with distill water until the pH became 7. This fiber was called pineapple leaf

microfiber (PALMF)

Silanizaton of pineapple leaf fiber

Bis-[γ-(triethoxysilyl)-propyl]-tetrasulfide (silane-69) was grafted onto the surface of

PALMF. A predetermined amount of silane-69 (3% wt. of fiber) was dissolved in absolute

ethanol (analytical grade) and stirred for 1 hour. Then PALMF was added into the silane

solution keeping the ratio of fiber and the solution at 1 g to 40 ml. The fiber was left in the

silane-69 solution for 72 hr. After that the solvent was removed by drying in a hot air oven at

80°C for 24 hr.

2.2 Fiber and NR masterbatch preparation

The masterbatch of PALMF-NR was prepared at 30% of fiber using a small scale

two-roll mill. First NR was masticated for 16 minutes. Then, PALMF was slowly added and

mixing continued for 12 minutes. Finally, it was taken from the mill.

5
2.3 Composite preparation

Compounds were prepared using a small scale two-roll mill. Rotor speed and nip gap

were kept the same for all formulations at 50 rpm and 0.25 mm, respectively. The sequence

of mixing is shown in Table 1. The Kevlar and pineapple leaf fiber in masterbatch forms

were used to achieve the final fiber content of 2, 5, and 10 phr.

2.4 Characterization

X-ray photoelectron spectroscopy (XPS)

The pineapple leaf fiber surface was analyzed with a VG Scienta SES2002

spectrometer equipped with a monochromatized Al(Kα) X-ray source (1486.6 eV), a

hemispherical analyzer and an electron gun to compensate the charging effect. The high-

resolution spectra and wide scan were recorded with pass energis of 100 eV and 500 eV,

respectively. The analyzed zone had a surface of 24 mm2. The decomposition of the spectra

into different components was performed with Gaussian-Lorentzian line shape, after having

subtracted a Shirley-type background. The mean composition of the sample surface

expressed in atomic percentages was determined using integrated peak areas of each

component taking into account the transmission factor of the spectrometer, the mean free

path, and the sensitivity factors of each atom. Wide-scan and high resolution scan were used.

The sample was measured under high vacuum. The x-rays can penetrate the surface of

sample to a depth of around 10 nm.

Dynamic mechanical analysis (DMA)

Rubber composites were characterized with a viscoanalyzer (METRAVIB VA 4000).

The specimens were cut into 30 x 10 x 1 mm3. Experiments were performed with fixed

gauge length of 20 mm. First, the specimen was set up in the closed chamber with controlled

6
temperature. The condition was 1% pre-strained, 0.5% dynamic strain, and 10 Hz frequency.

The sample was measured in tension mode. The temperature was decreased from room

temperature to -90°C with liquid nitrogen at the scan rate 10°C/min. The specimen was held

at -90°C for 15 min. Then, the temperature was increased from -90°C to 100°C at the rate of

2°C/min.

Scanning electron microscope (SEM)

Morphologies of fiber and rubber composites were observed with FEI Quanta 400

scanning electron microscope at different magnifications. Samples were stuck to carbon tape

and coated with gold before observation.

3. Results

3.1 PALMF

PALF was defibrillated and transformed into pineapple leaf microfiber (PALMF) as

described above. The size, shape and the surface of PALMF are shown in Figure 2.

3.2 XPS of Silanized PALMF

Silanization was used to modify the PALMF surface in order to promote the

interfacial adhesion with the NR matrix. XPS was used to confirm the presence of silane-69

on the PALMF surface. Figure 3 displays wide scan and high-resolution spectra of the

controlled PALMF and Si69 treated PALMF (Si69PALMF). The spectra show the presence

of sulfur and silicon atoms on the surface and their contents are 2.42 and 2.57, respectively.

The analysis of XPS spectra of Si69PALMF for the C1s peak indicates the amount of C-C

increases from 26.43% to 47.39%, while that of C-OR and O-C-O decrease from 33.58% to

20.5% and 6.72% to 2.84%, respectively (Table 2). The O/C ratio for Si69PALMF decreases

7
from 0.50 to 0.27. These results indicate that silane-69 was successfully grafted onto the

surface of PALMF.

3.3 Dynamic Mechanical Analysis

Mechanical properties of NR reinforced with 2, 5, and 10 phr of pineapple leaf

microfiber and Kevlar were determined at very low strain. Figure 4 shows the storage

modulus and tan delta over a range of temperature of the composites. At low temperatures all

composites exhibit high storage modulus and the effects of fiber and fiber content are

relatively small. As the temperature is increased, the modulus remains roughly constant until

about -50 °C when the modulus drops significantly. It is in this region that the effect of both

fibers in improving modulus is more obvious and the effect becomes greater with increased

fiber loading. As temperature is increased further, the modulus remains almost constant

again. Tan delta of the composites shows a broad peak at a temperature of about -44 °C and

the magnitude of tan delta at the peak decreases in line with increasing fiber loading.

However, the difference in the effect of Kevlar and PALMF is difficult to observe in these

graphs. In order to illustrate the difference, the moduli and tan delta values of the composites

at specific temperatures were extracted and are shown in Figure 5. At temperature lower than

the glass transition temperature (Tg), -75 °C, storage moduli of composites with low fiber

content of 2 phr are very similar. As fiber content increases, it becomes apparent that Kevlar

fiber has a slightly greater effect than does PALMF. As the temperature is increased to 25 °C

and 60 °C, the effect of both fibers remains similar at low fiber contents. Only at a fiber

content of 10 phr does PALMF demonstrate a much greater effect that does Kevlar. Also, the

tan delta peaks of the PALMF composites display correspondingly lower values.

8
3.4 Fracture surfaces

Figure 6 displays the cryogenic fracture surfaces of the composites in the plane

perpendicular to the fiber orientation direction. The spotty images clearly indicate uniaxial

orientation and also good dispersion of the fiber. High resolution images also indicate good

wetting of the matrix on the fibers.

4. Discussion

Pineapple leaf microfiber (PALMF) was prepared from PALF by alkali treatment and

mechanical force. Generally, alkali solution removes cementing material that holds

elementary microfibers together to form bundle. So the treatment carried out here provides

two simultaneous effects, i.e, breaking down bundles and cleaning the fiber surfaces for

effective later silanization [24]. The PALMF shows a smooth surface similar to that of alkali

treated PALF [22]. However, the cementing material cannot be completely removed as seen

from the XPS results in Figure 3. The C/O ratio of the PALMF surface is 0.50. This is closer

to the C/O ratio of lignin at 0.33 [25, 26] than that of cellulose at 0.83. Although lignin can

be completely removed, this would significantly weaken the microfiber [26, 27]. Despite the

remaining materials on the fiber surface, some grafting of silane is confirmed using XPS.

This layer is very thin and could not be observed with using SEM. That is similar to what is

observed on silane-treated sisal [28].

All fibers are well dispersed in the matrix rubber as seen in the SEM images in Figure

6. The images also show that the fiber can be uniaxially oriented to a very high degree by

using a two-roll mill.

The objective of the research was to compare the mechanical properties of the natural

and synthetic filler composites. At low fiber content, aramid fiber performs slightly better

than PALMFs in reinforcing rubbers. As the fiber content is increased to 10 phr, both types

9
of PALMFs overtake aramid fiber in their reinforcing effect. This clearly demonstrates the

potential of PALMF as a high-performance natural reinforcing fiber.

This fact raises the question as to how it could occur given that PALMF has a lower

modulus than does aramid fiber. The modulus of an uniaxial short fiber composite can be

calculated from different models. A useful one is the shear lag model originally proposed by

Cox [29]. This can be written in the followings simple form.

= + 1− (1)

/
= 1 −
/
(2)

= $
(3)
! "# '
%&

Where E1 is elastic modulus of composite, Em is modulus of matrix, Ef is modulus of fiber,

Vf is fiber volume fraction, d is diameter of fiber, l is fiber length, Gm is shear modulus of

matrix, is shear lag factor or modulus adjustment variable for short fiber, is a constant

value that depends on shape and orientation of fiber. It is assumed that the composite has

uniaxial orientation of cylindrical fibers with square arrangement.

Although there are a few fiber composite models that differ in detail, all depend on

the modulus of the reinforcing fiber and its aspect ratio. The latter contributes to the term ηl

and the higher the aspect ratio the closer it gets to unity. Thus, for reinforcing fiber with

lower modulus, one can compensate by using longer fibers.

Aramid fiber can have a modulus from about 64 GPa for as-spun fiber and above that

for different heat-treated grades. That of PALF has been reported to be 34.5-82.5 GPa [13].

We have found that, for the variety of pineapple used in this study, when in the form of

bundle it has a modulus of about 50 GPa [22]. It is likely that PALMF in its elementary fiber

form would have a greater modulus, The PALMF used in this work is 6 mm long and its

10
diameter is around 3 µm, giving it a very high aspect ratio of 2000. Aramid fiber, according

to product specification, is only 1-2 mm long and about 10 µm in diameter yielding an aspect

ratio in the range of 100-200. However, fiber breakage did occur during the mixing and the

size of the fiber can be determined by considering the toluene extracted fibers from the two

mixtures. The images of fibers extracted from uncured rubber compounds are shown in

Figure 7 and it is still confirmed that PALF is longer than aramid fiber. Such a large

difference in aspect ratio could cause the ηl to increase from 0.22 for aramid fiber to 0.74 for

PALMF. This is more than 70.5 % increase. It is in this way that PALMFs show a greater

effect on the rubber composite modulus than does the aramid fiber.

We turn now to look at the effect of temperature on these mechanical properties. At

very low temperatures, when the matrix is glassy and its modulus is closer to the reinforcing

fiber, the modulus of all composites are very close to each other being less than 5% apart. As

the temperature increases, the matrix becomes softer and rubbery and matrix deformation

makes a significant contribution to the total deformation of the composite. Stress transfer

from matrix to the reinforcing fiber occurs through better friction and traction at the interface.

In this state of large matrix deformation and high traction, the advantage of very high aspect

ratio PALMF becomes clearer.

There are also other contributing factors. Generally, mechanical modelling starts with

a representative elementary cell and properties derived from that single isolated cell. It has

been shown that, by using finite element method, the way these isolated cells pack or arrange

together affects the modulus of the composite. Hexagonal arrangement of the cells with a

regular overlapping of fiber ends will provide a higher modulus than does the arrangement

without overlapping [30]. The difference becomes greater with increasing fiber volume

fraction. So PALMF, with an aspect ratio much greater than aramid will have more overlap

and greater reinforcement.

11
Treatment of PALMF with silane-69 improves not only the adhesion between fiber

and NR matrix, but also the filler dispersion. It does this by reducing the polarity difference

between the reinforcing fiber and the matrix rubber. The silane-69 treatment of the fiber

allows the formation of chemical bonds bridging the fiber surface with the rubber [31], thus

improving the interfacial adhesion. As a result, the overall modulus of so treatd composite is

greater than that of unmodified PALMF-NR composites. Such an effect of silane treated

fibers on the properties of rubber has been reported elsewhere [32].

5. Conclusions

Silane-69 can be successfully grafted onto the surface of PALMF. Both untreated

PALMF and silane-69 treated fiber exhibit reinforcing efficiency comparable to that of

Kevlar fiber. Silane treatment yields fibers with a slight improvement over the untreated

counterparts. At high fiber content, both types of PALMF exhibit greater reinforcing

efficiency than does Kevlar fiber. The reinforcing mechanism in short fiber reinforced rubber

composites is different at low and high fiber contents. At low fiber content the modulus of

the reinforcing fiber dominates. At high fiber content the fiber aspect ratio becomes more

important. So naturally sourced PALMF has very high potential as a high performance

reinforcement for rubber matrix composites.

Acknowledgements

The authors gratefully acknowledge partial financial support of the project by the

Center of Excellence for Innovation in Chemistry (PERCH-CIC), Office of the Higher

Education Commission, Ministry of Education. We acknowledge also the grant of Thai

scholarships from Center of Excellence for Innovation in Chemistry (PERCH-CIC) and

Franco-Thai scholarship from the French government provided for BS.

12
Data Availability

The raw/processed data required to reproduce these findings are available upon

request.

References

[1] What is NAFILean ?, Automot. Perform. Mater. https://www.apm-


planet.com/products/nafilean/what-is-nafilean/ (accessed August 8, 2019).
[2] S. Akhtar, A.K. Bhowmick, P.P. De, S.K. De, Tensile rupture of short fibre filled
thermoplastic elastomer, J. Mater. Sci. 21 (1986) 4179–4184. doi:10.1007/BF01106527.
[3] D.K. Setua, S.K. De, Effect of short fibres on critical cut length in tensile failure of
rubber vulcanizates, J. Mater. Sci. 20 (1985) 2653–2660. doi:10.1007/BF00556098.
[4] H. Ismail, M.R. Edyham, B. Wirjosentono, Bamboo fibre filled natural rubber
composites: The effects of filler loading and bonding agent, Polym. Test. 21 (2002)
139–144. doi:10.1016/S0142-9418(01)00060-5.
[5] V.G. Geethamma, K.T. Mathew, R. Lakshminarayanan, S. Thomas, Composite of short
coir fibres and natural rubber: Effect of chemical modification, loading and orientation
of fibre, Polymer. 39 (1998) 1483–1491. doi:10.1016/S0032-3861(97)00422-9.
[6] M. Jacob, S. Thomas, K.T. Varughese, Mechanical properties of sisal/oil palm hybrid
fiber reinforced natural rubber composites, Compos. Sci. Technol. 64 (2004) 955–965.
doi:10.1016/S0266-3538(03)00261-6.
[7] V.M. Murty, S.K. De, Shot jute fiber reinforced rubber composites, Rubber Chem.
Technol. 55 (1982) 287–308. doi:10.5254/1.3535879.
[8] M.J. John, B. Francis, K.T. Varughese, S. Thomas, Effect of chemical modification on
properties of hybrid fiber biocomposites, Compos. Part Appl. Sci. Manuf. 39 (2008)
352–363. doi:10.1016/j.compositesa.2007.10.002.
[9] H. Anuar, A. Zuraida, Improvement in mechanical properties of reinforced
thermoplastic elastomer composite with kenaf bast fibre, Compos. Part B Eng. 42
(2011) 462–465. doi:10.1016/j.compositesb.2010.12.013.
[10] N. Lopattananon, K. Panawarangkul, K. Sahakaro, B. Ellis, Performance of pineapple
leaf fiber-natural rubber composites: The effect of fiber surface treatments, J. Appl.
Polym. Sci. 102 (2006) 1974–1984. doi:10.1002/app.24584.

13
[11] S. Varghese, B. Kuriakose, S. Thomas, Stress relaxation in short sisal‐fiber‐reinforced
natural rubber composites, J. Appl. Polym. Sci. 53 (1994) 1051–1060.
doi:10.1002/app.1994.070530807.
[12] W. Wang, G. Huang, Characterisation and utilization of natural coconut fibres
composites, Mater. Des. 30 (2009) 2741–2744. doi:10.1016/j.matdes.2008.11.002.
[13] V.M. Murty, S.K. De, Effect of particulate fillers on short jute fiber‐reinforced natural
rubber composites, J. Appl. Polym. Sci. 27 (1982) 4611–4622.
doi:10.1002/app.1982.070271208.
[14] U. Wisittanawat, S. Thanawan, T. Amornsakchai, Mechanical properties of highly
aligned short pineapple leaf fiber reinforced - Nitrile rubber composite: Effect of fiber
content and Bonding Agent, Polym. Test. 35 (2014) 20–27.
doi:10.1016/j.polymertesting.2014.02.003.
[15] U. Wisittanawat, S. Thanawan, T. Amornsakchai, Remarkable improvement of failure
strain of preferentially aligned short pineapple leaf fiber reinforced nitrile rubber
composites with silica hybridization, Polym. Test. 38 (2014) 91–99.
doi:10.1016/j.polymertesting.2014.07.006.
[16] K. Prukkaewkanjana, S. Thanawan, T. Amornsakchai, High performance hybrid
reinforcement of nitrile rubber using short pineapple leaf fiber and carbon black, Polym.
Test. 45 (2015) 76–82. doi:10.1016/j.polymertesting.2015.05.004.
[17] K. Yantaboot, T. Amornsakchai, Effect of mastication time on the low strain properties
of short pineapple leaf fiber reinforced natural rubber composites, Polym. Test. 57
(2017) 31–37. doi:10.1016/j.polymertesting.2016.11.006.
[18] K. Yantaboot, T. Amornsakchai, Effect of preparation methods and carbon black
distribution on mechanical properties of short pineapple leaf fiber-carbon black
reinforced natural rubber hybrid composites, Polym. Test. 61 (2017) 223–228.
doi:10.1016/j.polymertesting.2017.05.026.
[19] P. Pittayavinai, S. Thanawan, T. Amornsakchai, Manipulation of mechanical properties
of short pineapple leaf fiber reinforced natural rubber composites through variations in
cross-link density and carbon black loading, Polym. Test. 54 (2016) 84–89.
doi:10.1016/j.polymertesting.2016.07.002.
[20] N. Hariwongsanupab, S. Thanawan, T. Amornsakchai, M.-F. Vallat, K. Mougin,
Improving the mechanical properties of short pineapple leaf fiber reinforced natural
rubber by blending with acrylonitrile butadiene rubber, Polym. Test. 57 (2017) 94–100.
doi:10.1016/j.polymertesting.2016.11.019.
14
[21] P. Pittayavinai, S. Thanawan, T. Amornsakchai, Comparative study of natural rubber
and acrylonitrile rubber reinforced with aligned short aramid fiber, Polym. Test. 64
(2017) 109–116. doi:10.1016/j.polymertesting.2017.09.033.
[22] B. Surajarusarn, P. Traiperm, T. Amornsakchai, Revisiting the Morphology,
Microstructure, and Properties of Cellulose Fiber from Pineapple Leaf so as to Expand
Its Utilization, Sains Malays. 48 (2019) 145–154. doi:10.17576/jsm-2019-4801-17.
[23] N. Kengkhetkit, T. Amornsakchai, Utilisation of pineapple leaf waste for plastic
reinforcement: 1. A novel extraction method for short pineapple leaf fiber, Ind. Crops
Prod. 40 (2012) 55–61. doi:10.1016/j.indcrop.2012.02.037.
[24] S. Kalia, B.S. Kaith, I. Kaur, Pretreatments of natural fibers and their application as
reinforcing material in polymer composites—A review, Polym. Eng. Sci. 49 (2009)
1253–1272. doi:10.1002/pen.21328.
[25] L.-S. Johansson, J.M. Campbell, K. Koljonen, P. Stenius, Evaluation of surface lignin
on cellulose fibers with XPS, Appl. Surf. Sci. 144–145 (1999) 92–95.
doi:10.1016/S0169-4332(98)00920-9.
[26] L.Q.N. Tran, C.A. Fuentes, C. Dupont-Gillain, A.W. Van Vuure, I. Verpoest, Wetting
analysis and surface characterisation of coir fibres used as reinforcement for composites,
Colloids Surf. Physicochem. Eng. Asp. 377 (2011) 251–260.
doi:10.1016/j.colsurfa.2011.01.023.
[27] P. Muensri, T. Kunanopparat, P. Menut, S. Siriwattanayotin, Effect of lignin removal on
the properties of coconut coir fiber/wheat gluten biocomposite, Compos. Part Appl. Sci.
Manuf. 42 (2011) 173–179. doi:10.1016/j.compositesa.2010.11.002.
[28] F. Zhou, G. Cheng, B. Jiang, Effect of silane treatment on microstructure of sisal fibers,
Appl. Surf. Sci. 292 (2014) 806–812. doi:10.1016/j.apsusc.2013.12.054.
[29] H.L. Cox, The elasticity and strength of paper and other fibrous materials, Br. J. Appl.
Phys. 3 (1952) 72–79. doi:10.1088/0508-3443/3/3/302.
[30] J.-M. Berthelot, A. Cupcic, K.A. Brou, Stress distribution and effective longitudinal
Young’s modulus of unidirectional discontinuous fibre composites, J. Compos. Mater.
27 (1993) 1391–1425. doi:10.1177/002199839302701404.
[31] A. Ansarifar, F. Saeed, S.O. Movahed, L. Wang, K.A. Yasin, S. Hameed, Using a
sulfur-bearing silane to improve rubber formulations for potential use in industrial
rubber articles, J. Adhes. Sci. Technol. 27 (2013) 371–384.
doi:10.1080/01694243.2012.705541.

15
[32] N. Lopattananon, D. Jitkalong, M. Seadan, Hybridized reinforcement of natural rubber
with silane-modified short cellulose fibers and silica, J. Appl. Polym. Sci. 120 (2011)
3242–3254. doi:10.1002/app.33374.

16
Table Captions

Table 1 Sequence of NR reinforced with aramid fiber mixing

Table 2 Surface elements, chemical state and atomic content from XPS analysis of

PALMFs.

17
Table 1 Sequence of NR reinforced with aramid fiber mixing

Step Ingredients Time (min)

1 NR 0-2

Masterbatch of aramid fiber or

2 Masterbatch of pineapple leaf 2-6

microfiber

3 ZnO and stearic acid 8-10

4 Sulfur and CBS 10-13

5 Sheet out 13-14

Total time 14

18
Table 2 Surface elements, chemical state and atomic content from XPS analysis of

PALMFs.

Building Energy (eV) % Atomic Content

Name Control Control


Si69PALMF Si69PALMF
PALMF PALMF

C-C, C-H 284.94 285.03 26.43 47.39

C1 286.58 (C-OR) 286.75 (C-O) 33.58 20.5

C1s C2 288.1 (C=O) 288.72 6.72 2.84

C-Si - 284.36 - 2.59

C-S - 284.66 - 1.23

O1 532.74 533.03 19.92 7.64

O1s O2 533.32 533.61 13.35 5.12

Si-O (silane) - 532.3 - 7.7

S 2p3/2 S-S - 164.18 - 1.21


S2p
S 2p3/2 C-S - 163.89 - 1.21

Si2p Si 2p3/2 Si(-O)3 - 102.21 - 2.57

19
Figure Captions

Figure 1 Steps in preparations of pineapple leaf microfiber (a) and rubber composites to

obtain unidirectional fiber alignment (b).

Figure 2 SEM micrograph of PALMF (a and b) and Si69-PALMF (c and d) at low and high

magnification.

Figure 3 XPS wide scan and high resolution spectra of C1s of PALMF (top) and silanized

PALMF (bottom).

Figure 4 Elastic moduli (Eꞌ) and tan δ of natural rubber and its composites containing

different amounts and types of fibers.

Figure 5 Elastic moduli of natural rubber and its composites containing different amounts

and types of fibers at varying temperature.

Figure 6 Cryogenic fracture surfaces of different composites displaying good fiber

distribution at low and high magnification. Aramid fiber(a and d), PALMF (b and

e) and Si69-PALMF (c and f) reinforced NR composites.

Figure 7 PALF (left) and aramid fiber (right) extracted from their respective uncured rubber

compounds.

1
(a)

(b)

Figure 1 Steps in preparations of pineapple leaf microfiber (a) and rubber composites to

obtain unidirectional fiber alignment (b).

2
(a) (b)

200.0 µm 20.0 µm

(c) (d)

200.0 µm 20.0 µm

Figure 2 SEM micrograph of PALMF (a and b) and Si69-PALMF (c and d) at low and high

magnification.

3
14000
PALMF O1s Wide scan PALMF
12000
C-OR

10000

C-C, C-H

Intensity
Intensity

8000
C1s
6000

4000 C=O

2000

0
1000 800 600 400 200 0 292 290 288 286 284 282 280

Binding energy (eV) Binding energy (eV)

0
6.0x10
Si69PALMF Si69PALMF
C1s
O1s Wide scan
C-C, C-H
0
4.0x10
Intensity
Intensity

0
2.0x10
C-OR
S2p
Si2s C2
S2s Si2p C-Si C-S
0.0
1000 800 600 400 200 0 292 290 288 286 284 282 280

Binding energy (eV) Binding energy (eV)

Figure 3 XPS wide scan and high resolution spectra of C1s of PALMF (top) and silanized

PALMF (bottom).

4
11
10 5.0
NR
4.5
10 K2NR
10 4.0 PALMF2
3.5 Si69PALMF2
9
10
3.0

Tan delta
E' (Pa)

8
10 2.5
2.0
7
10 1.5
6 1.0
10
0.5
5
10 0.0
-50 0 50 -50 0 50
11 5.0
10
NR
4.5
10 K5NR
10 4.0 PALMF5
3.5 Si69PALMF5
9
10
3.0
Tan delta
E' (Pa)

8 2.5
10
2.0
7
10 1.5
6 1.0
10
0.5
5 0.0
10
-50 0 50 -50 0 50
11
10 5.0
NR
4.5
10 K10NR
10 4.0 PALMF10
9 3.5 Si69PALMF10
10
3.0
Tan delta
E' (Pa)

8
10 2.5
2.0
7
10
1.5
6 1.0
10
0.5
5
10 0.0
-50 0 50 -50 0 50
Temperature (°C) Temperature (°C)

Figure 4 Elastic moduli (Eꞌ) and tan δ of natural rubber and its composites containing

different amounts and types of fibers.

5
6000
NR -75 οC
K
5000 PALMF
Si69PALMF
4000
E' (MPa)
3000

2000

1000

0
0 2 5 10
40
25 οC

30
E' (MPa)

20

10

0
0 2 5 10
30
60 οC

20
E' (MPa)

10

0
0 2 5 10
Fiber content (phr)

Figure 5 Elastic moduli of natural rubber and its composites containing different amounts

and types of fibers at varying temperature.

6
(a) (b) (c)

200.0 µm 200.0 µm 200.0 µm

(d) (e) (f)

10.0 µm 10.0 µm 10.0 µm

Figure 6 Cryogenic fracture surfaces of different composites displaying good fiber

distribution at low and high magnification. Aramid fiber(a and d), PALMF (b and

e) and Si69-PALMF (c and f) reinforced NR composites.

7
1 mm 1 mm

Figure 7 PALF (left) and aramid fiber (right) extracted from their respective uncured rubber

compounds.

8
Highlights

• Reinforcement efficiency of short pineapple leaf microfiler and aramid fiber in natural

rubber are compared.

• Comparison is based on low strain behaviour from dynamic mechanical analysis.

• At low fiber content of 2 phr, aramid fiber provides slightly higher reinforcing effect

than does pineapple leaf microfiber.

• At high fiber content of 10 phr, pineapple leaf microfiber provides significantly

higher reinforcing effect than does aramid fiber.


Author contributions
Name CRediT roles
Budsaraporn Surajarusarn Investigation, Writing - Original Draft
Samar Hajjar-Garreau Methodology, Investigation
Gautier Schrodj Methodology, Investigation
Karine Mougin Validation, Review
Taweechai Amornsakchai Conceptualization, Writing - Review & Editing
Declaration of interests

☒ The authors declare that they have no known competing financial interests or personal relationships
that could have appeared to influence the work reported in this paper.

☐The authors declare the following financial interests/personal relationships which may be considered
as potential competing interests:

You might also like