You are on page 1of 14

Composites: Part A 41 (2010) 806819

Contents lists available at ScienceDirect

Composites: Part A
journal homepage: www.elsevier.com/locate/compositesa

Review

Silane coupling agents used for natural ber/polymer composites: A review


Yanjun Xie a,b,*, Callum A.S. Hill b, Zefang Xiao a, Holger Militz a, Carsten Mai a
a
b

Wood Biology and Wood Products, Burckhardt-Institute, Georg August University of Gttingen, Bsgenweg 4, D37077 Gttingen, Germany
Centre for Timber Engineering, School of Engineering and the Built Environment, Edinburgh Napier University, 10 Colinton Road, EH10 5DT Edinburgh, United Kingdom

a r t i c l e

i n f o

Article history:
Received 9 November 2009
Received in revised form 4 March 2010
Accepted 4 March 2010

Keywords:
A. Silane coupling agents
A. Fibres
B. Fiber/matrix bond
B. Environmental degradation

a b s t r a c t
Natural ber reinforced polymer composites (NFPCs) provide the customers with more alternatives in the
material market due to their unique advantages. Poor bermatrix interfacial adhesion may, however,
negatively affect the physical and mechanical properties of the resulting composites due to the surface
incompatibility between hydrophilic natural bers and non-polar polymers (thermoplastics and thermosets). A variety of silanes (mostly trialkoxysilanes) have been applied as coupling agents in the NFPCs to
promote interfacial adhesion and improve the properties of composites. This paper reviews the recent
progress in using silane coupling agents for NFPCs, summarizes the effective silane structures from the
silane family, claries the interaction mechanisms between natural bers and polymer matrices, and presents the effects of silane treatments on the mechanical and outdoor performance of the resulting
composites.
Crown Copyright 2010 Published by Elsevier Ltd. All rights reserved.

1. Introduction and background


Natural ber reinforced polymer composites (NFPCs), as an
important branch in the eld of composite materials, have been
studied for decades [111]. Natural bers have different origins
such as wood, pulp, cotton, bark, nut shells, bagasse, corncobs,
bamboo, cereal straw, and vegetable (e.g., ax, jute, hemp, sisal,
and ramie) [1013]. These bers are mainly made of cellulose,
hemicelluloses, lignin and pectins, with a small quantity of
extractives. The ber constituents vary depending on their origi-

Abbreviations: ABI, acidbase interaction; APS, c-aminopropyl triethoxy silane;


ASE, anti-swelling efciency; BDMA, benzyl dimethylamine; BPO, benzoyl peroxide; CTMP, chemithermomechanical pulp; DCUP, dicumyl peroxide; DCS, dichlorodiethylsilane; DGEBA, diglycidyl ethers of bisphenol A; DMA, dynamic
mechanical analysis; DMVS, dichloro methyl vinyl silane; DSC, differential scanning
calorimetry; EMC, equilibrium moisture content; FTIR, Fourier transform infrared
spectroscopy; GPS, c-glycidoxypropyltrimethoxy silane; HDS, hexadecyltrimethoxy
silane; IPN, interpenetrating polymer network; ISS, interfacial shear strength; LDPE,
low density polyethylene; MAPP, maleated polypropylene; MEKP, methyl ethyl
ketone peroxide; MMS, methacryloxymethyltrimethoxy silane; MPS, c-methacryloxypropyl trimethoxy silane; MRPS, c-mercaptopropyltrimethoxy silane; NFPCs,
natural ber reinforced polymer composites; PALF, pineapple leaf ber; PAPS, cphenyl-aminopropyltrimethoxy silane; PE, polyethylene; PP, polypropylene; PS,
polystyrene; PVC, polyvinyl chloride; SEC, size exclusion chromatography; TAS, cdiethylenetriaminopropyl trimethoxy silane; TGA, thermogravimetric analysis; UP,
unsaturated polyester; VSPP, vinyltrimethoxysilane grafted polypropylene; VTS,
vinyltrimethoxy silane; XPS, X-ray photoelectron spectroscopy.
* Corresponding author at: Centre for Timber Engineering, School of Engineering
and the Built Environment, Edinburgh Napier University, 10 Colinton Road, EH10
5DT Edinburgh, United Kingdom. Tel.: +44 131 4552494; fax: +44 131 4552239.
E-mail address: y.xie@napier.ac.uk (Y. Xie).

nation. Compared with conventional inorganic llers such as glass


ber and carbon bers, natural bers provide many advantages:
(1) abundance and therefore low cost, (2) biodegradability, (3)
exibility during processing and less resulting machine wear,
(4) minimal health hazards, (5) low density, (6) desirable ber aspect ratio, and (7) relatively high tensile and exural modulus.
Incorporating the tough and light-weight natural bers into polymer (thermoplastic and thermoset) matrices produces composites
with a high specic stiffness and strength [14]. The renewable
and biodegradable characteristics of natural bers facilitate their
ultimate disposal by composting or incineration, options not possible with most industrial bers. The bers also contain sequestered atmospheric carbon dioxide in their structure and are
invariably of lower embodied energy compared to industrially
produced glass bers.
Although natural bers can offer the resulting composites many
advantages, the usually polar bers have inherently low compatibility with non-polar polymer matrices, especially hydrocarbon
matrices such as polypropylene (PP) and polyethylene (PE)
[15,16]. The incompatibility may cause problems in the composite
processing and material properties. Hydrogen bonds may form between the hydrophilic bers, and thus the bers tend to agglomerate into bundles and unevenly distribute throughout the non-polar
polymer matrix during compounding processing [17,18]. There is
also insufcient wetting of bers by the non-polar polymer matrices, resulting in weak interfacial adhesion. As a result, the stress
transfer efciency from the matrix to the reinforcing bers is reduced. The incompatibility may not be an issue when using polar
polymers such as unsaturated polyester (UP) and epoxy resin as

1359-835X/$ - see front matter Crown Copyright 2010 Published by Elsevier Ltd. All rights reserved.
doi:10.1016/j.compositesa.2010.03.005

Y. Xie et al. / Composites: Part A 41 (2010) 806819

matrices; however, the resulting composites, similar to the composites with non-polar matrices, will be susceptible to moisture
deterioration and fungal damage during outdoor service. The moisture absorption of the natural bers may cause dimensional
changes of the resulting composites and weaken the interfacial
adhesion [19,20]. Mould and decay fungi may also grow on/in
the composites, although more slowly than in the bers alone,
when they are utilized in the long-term under wet conditions
[21]. In addition, natural bers are of limited thermal stability
and, therefore, thermal degradation may take place during composite processing at a high temperature, especially in the cases of
thermal extrusion and hot compression processes.
By this token, treatment of natural bers is benecial in order to
improve the water resistance of bers, enhance the wettability of
the natural ber surface by polymers (mainly non-polar polymers)
and promote interfacial adhesion. The performance of bers is critical to obtain the improved physical and mechanical properties of
the resulting composites. Physical treatments (e.g. electronic discharge in the different media such as plasma and corona technologies [2225]) may create a hydrophilic or hydrophobic ber
surface by changing the surface energy to consequently increase
the compatibility of the treated ber with the polymer matrices.
These surface treatments only modify a very shallow surface of cell
walls and thus do not change the hygroscopic characteristics of bers. Chemical modication provides the means of permanently
altering the nature of ber cell walls [26]: by grafting polymers
onto the bers [27,28], crosslinking of the ber cell walls [29], or
by using coupling agents [4]. These modifying strategies have been
generally reviewed recently [26,30]. The chemical modication
may make the ber cell walls more dimensionally stable, reduce
water sorption, or increase resistance against fungal decay, but
there may be an associated reduced dynamic strength such as impact strength due to embrittlement. A coupling agent is a chemical
that functions at the interface to create a chemical bridge between
the reinforcement and matrix. It improves the interfacial adhesion
when one end of the molecule is tethered to the reinforcement surface and the functionality at the other end reacts with the polymer
phase. Extensively used coupling agents for NFPCs are copolymers
containing maleic anhydride such as maleated polypropylene
(MAPP) or maleated polyethylene (MAPE) [18,3133]. The anhydride groups of the copolymers may react with the surface hydroxyl groups of natural bers forming ester bonds whilst the other
end of copolymer entangles with the polymer matrix due to their
similar polarities [34]. Isocyanates have also been reported as the
coupling agents used in NFPCs. Urethane links can be formed between the isocyanate functionality and the hydroxyl group of natural bers [35,36], consequently blocking these hygroscopic
hydroxyl sites [37].
Silanes are recognized as efcient coupling agents extensively
used in composites and adhesive formulations [38]. They have
been successfully applied in inorganic ller reinforced polymer
composites such as glass ber reinforced polymer composites
[e.g. 39,40] and mineral lled polymer composites [e.g. 41,42]. Silanes are also adhesion promoters in many adhesive formulations
or are used as substrate primers, giving stronger adhesion [43]. The
bifunctional structures of silanes have also been of interest in
applying them for natural ber/polymer composites, since both
glass bers and natural bers bear reactive hydroxyl groups, and
extensive researches have accordingly been carried out to screen
the varied silane structures for NFPC production.
The aim of this paper is to review the recent progress in using
silane coupling agents for the production of natural ber reinforced
polymer composites. The effects of treatments of natural bers
with silanes on the properties of bers and the resulting composites are discussed and the interaction mechanisms between phases
claried.

807

2. Silane structures and hydrolysis processes


2.1. Silane structures
To effectively couple the natural bers and polymer matrices,
the silane molecule should have bifunctional groups which may
respectively react with the two phases thereby forming a bridge
in between them. Silane coupling agents have a generic chemical
structure R(4n)ASiA(R0 X)n (n = 1,2) where R is alkoxy, X represents an organofunctionality, and R0 is an alkyl bridge (or alkyl
spacer) connecting the silicon atom and the organofunctionality.
In the past decades, various silane structures have been tested
for coupling of inorganic reinforcements such as glass ber and
organic polymer matrices. The structures used to couple the natural bers and polymer matrices are relatively limited. Most of
the established silanes used for NFPCs are trialkoxysilanes. The
organofunctionality of the silane interacts with the polymer
matrices with their interaction modes depending on the functionalitys reactivity or compatibility towards the polymer. A nonreactive alkyl group of the silane may increase the compatibility
with non-polar matrix due to their similar polarities; however,
the reactive organofunctionality may covalently bond with as
well as being physically compatible with the polymer matrices.
These organofunctionalities of silanes are typically amino, mercapto, glycidoxy, vinyl, or methacryloxy groups. The most reported silanes and their applied target polymer matrices are
listed in Table 1.
With regard to these silanes shown in Table 1, aminosilanes,
especially c-aminopropyltriethoxysilane (APS), are most extensively reported in the literature as coupling agents between natural
bers and thermoplastics or thermosets. Vinyl- and acryl-silanes
are coupling agents that are able to establish covalent bonds with
polymeric matrices in the presence of peroxide initiators. Methacrylatefunctional silanes can display high levels of reactivity with
unsaturated polyester matrices [44], whilst azidosilanes can efciently couple inorganic llers with thermoplastic matrices
[62,63]; however, there have been few reports of their use in natural ber reinforced thermoplastic composites. The application and
effects of these typical silanes on the NFPCs properties will be presented below.

2.2. Hydrolysis processes of silanes


The alkoxysilanes have been demonstrated to be able to directly
react with ASiAOH groups of silica thereby forming ASiAOASiA
bonds [e.g. 43,67,68] without any requirement of prehydrolysis.
However, silanes do not undergo the same reaction with the hydroxyl groups of cellulosic bers even at high temperature [69].
This has been attributed to lower acidity of cellulosic hydroxyl
groups compared with silanol [69]. In addition, cellulose is generally unreactive to many chemicals and the OH groups of the microbrils have very low accessibility. Based on the fact, an optional
strategy is to activate the alkoxysilane by hydrolyzing the alkoxy
groups off thereby forming the more reactive silanol groups. As a
result, the silanol may react with the hydroxyl groups of bers or
condense themselves on the surfaces of bers and/or in the cell
walls forming macromolecular network. Although the formed
ASiAOACA bonds are eventually not stable towards hydrolysis,
blocking the hydroxyl groups (reversible to hydrolysis) and the formation of macromolecular network (permanent) under heating
condition facilitate an enhancement of interfacial adhesion of treated bers and polymer matrices and of the properties of the resulting composites. To hydrolyze the alkoxy groups off, participation of
water is essential. Even though the natural bers under room
condition contain bound water which may act to hydrolyze the

808

Y. Xie et al. / Composites: Part A 41 (2010) 806819

Table 1
Silanes used for the natural ber/polymer composites: chemical structures, organofunctionalities and target polymer matrices.

a
b

Structure

Functionality

Abbreviation

Target matrix

References

(RO)3SiA(CH2)3ANH2a

Amino

APS

Epoxy
Polyethylene
Butyl rubber
Polyacrylate
PVC

[19,4447]

(RO)3SiACH@CH2

Vinyl

VTS

Polyethylene
Polypropylene
Polyacrylate

[44,4852]

(RO)3SiA(CH2)3AOOC(CH3)C@CH2

Methacryl

MPS

Polyethylene
Polyester

[20,44,53,54]

(RO)3SiA(CH2)3ASH

Mercapto

MRPS

Natural rubber
PVC

[53,5557]

(RO)3SiA(CH2)3AOACH2CHCH2O

Glycidoxy

GPS

Epoxy
Butyl rubber
Polysulde

[19,45,58,59]

R2ASiACl2

Chlorine

DCS

Polyethylene
PVC

[47,60]

VTS grafted plastics

Vinyl

VSPP
VSPE

Polypropylene
Polyethylene

[52,61]

(RO)3ASiAR00 AN3b

Azide

ATS

Polypropylene
Polyethylene
Polystyrene

[6264]

(RO)3SiA(CH2)15CH3

Alkyl

HDS

Polyethylene
Natural rubber

[53,65,66]

R = Amethyl or ethyl.
R00 = AC6H4ASO2A.

silanes; however, additional water is required to achieve a complete hydrolysis of silanes [7072].

2.2.1. Effect of silane structures on the hydrolysis


Alkoxy groups associated with silane coupling agents prior to
utilization can be hydrolyzed off thereby liberating the corresponding alcohols in the presence of water and generating reactive
silanol groups. The alkoxy groups are usually ethoxy or methoxy.
Under the same hydrolytic conditions the methoxy groups of trimethoxysilane hydrolyzes more rapidly than the ethoxy groups
of triethoxysilane [73]. The hydrolysis of trimethoxysilane producing methanol may be more environmentally problematic than the
triethoxysilane releasing ethanol because methanol is more toxic
than ethanol. The number of alkoxy groups will determine the
amount of water used to fully hydrolyze them and inuence the
adhesion between silanes and ller. Di- and tri-alkoxy silanes produce stronger adhesion strength than mono-alkoxy silanes since
they form more binding sites after they are hydrolyzed [74]. If it
is desired to deliver an alkoxy functional silane to the interior of
the ber this can be accomplished by dissolving the silane in the
appropriate alcohol (e.g. ethanol for an ethoxy silane). Hydrolysis
can then be achieved by subsequent water treatment.
The length of the alkyl spacer between the silane organofunctionality and the silicon atom also inuences the hydrolysis rate
of the silane (Fig. 1a and b). For example, a-methacryloxymethyltrimethoxy silane (MMS, a-silane) hydrolyzes 20 times faster than
c-methacryloxypropyltrimethoxy silane (MPS, c-silane) in an acetone/water mixture at pH 4 [75]. This has been attributed to the socalled a-effect of silanes. The shorter alkyl spacer causes higher
reactivity of alkoxy groups due to the stronger electron interaction
between the functionality (X) and the silicon atom. Such effects
may be explained using a-aminoalkylsilanes as an example, where
back-bonding of the nitrogens free electron pair to the silicon
atom weakens the SiAO bonds thereby causing a higher reactivity
of alkoxy group (Fig. 1c).

2.2.2. Effect of hydrolytic conditions on the hydrolysis of silane


In the presence of water, the silane will undergo a hydrolysis
process. The alkoxysilanes are hydrolyzed thereby forming alkoxysilanol mono- and di-, and then eventually silane tri-ols. To fully
hydrolyze one mole of trialkoxysilane, at least three moles of water
is needed. Once the silanol groups form in the solution, the condensation between silanol groups also starts, therefore generating
a siloxane (ASiAOASiA) polymer network in the solution. Thus,
the sol evolves towards the formation of a gel-like diphasic system
containing both a liquid phase and solid phase whose morphologies range from discrete particles to continuous polymer networks.
The competition of alkoxy hydrolysis and silanol condensation are

OR

(a)

RO

Si

CH2

OR
OR

(b)

RO

Si

(CH2 )3

OR
CH2

R1(O)

(c)

R1O
R1(O)

Si

R2
R2

Fig. 1. Different length of the bridge connecting the silicon atom and functionalities: a-silane having a methylene spacer (a), c-silane bearing a propylene spacer
(b), and the schematic diagram of the a-effect in the aminomethylene silane (c)
[75].

809

Y. Xie et al. / Composites: Part A 41 (2010) 806819

affected by the hydrolysis condition used such as the solvent, temperature, pH, concentration of silanes [7678]. Under optimum
conditions hydrolysis may be accelerated, but condensation of silanols is inhibited to maintain a stable intermediary structures such
as silanol monomer or dimers [76,79,80].
The dynamics of the hydrolysis and ensuing condensation of
several trimethoxysilanes bearing different chain length of alkyl
functionalities (methyl, ethyl, propyl, or butyl) have been investigated in wateracetone mixtures by quantifying the produced
methanol using internal reection Fourier transform infrared spectroscopy (FTIR) combined with statistical regression. The results
show that the relative rates of the consecutive hydrolysis reactions
and the rate of condensation to form siloxane bonds depend on the
silane organofunctionalities, and this FTIR technique permits the
tailoring of the composition of the alkoxysilane compounds in
solution to obtain the desired rate and degree of hydrolysis of alkoxy groups, and the extent of siloxane bond formation [82]. The
hydrolysis rates of APS, c-diethylenetriaminopropyl trimethoxy silane (TAS) and MPS were established using 1H NMR by measuring
the evolution of liberated free alcohol, and it showed that the rate
increased in the order MPS < APS < TAS. The formation of the silanol groups was followed by their condensation to generate oligomeric structures. Whereas APS and MPS only gave soluble
products, colloidal particles by contrast precipitated in the medium when the TAS was hydrolyzed [83,84]. At alkaline pH, the alkoxy groups of APS were hydrolyzed almost immediately in water
and the resulting aminosilanols formed a cyclic structure with an
internal hydrogen-bonded ring (Fig. 2). Bonding of the silanol
hydrogen makes the silanol stable against condensation in the concentrated aqueous solutions. As a result, there was no insoluble gel
in an alkaline aqueous solution of the aminoalkylsilanes after several months of storage [43]. In the pH range of 36, aminosilanol
groups in the solution were relatively more stable, existing as zwitterionic structures [43]. In contrast, MPS hydrolyzes much slower
and requires the use of an amine as catalyst in the water/solvent
solution [83].
Generally, under acid-catalyzed conditions, the hydrolysis rate
of silanes forming silanol groups is greater than the condensation
rate of the ensuing silanols forming siloxane bonds [72,79]. These
competing reactions of hydrolysis and condensation in an ethanol/water (80/20, w/w) solution at different pH values were ascertained in situ using 1H, 13C, and 29Si NMR spectroscopy [72]. After
hydrolysis of the silane, the ensuing silanol groups gradually selfcondensed into the structures of dimer (T1) (Fig. 3a), linear siloxane
(T2) (Fig. 3b), and ultimately rigidly three-dimensional polysiloxane cage structures (T3) (Fig. 3c). In an acidic solution, the hydrolysis rate of MPS was accelerated and the self-condensation rate was
reduced compared to those in the alkaline solution. The entities
with T3 units appeared after more than 1 month under acidic conditions; however, they appeared after only 6 h in the alkaline solution. The formation of a T3 structure in the treating solution must
be inhibited since this may reduce the number of silanol groups
adsorbed to bers. The silanol condensation will also make the cell
wall penetration by silane impossible due to the increased molecular size.

CH2CH2

CH2

Si
O

OH

NH2

Fig. 2. A proposed cyclic hydrogen-bonded amine structure in the hydrolytic APS


solution [81].

R'

OR

Si

Si

OR

RO

(a)

RO

R'

T1

structure
dimer or chain end
R'

(b)

OR

RO

R'
R'

Si
RO

Si
O

OR
Si

OR

T2 structure
linear link
R'

R'

R'

Si

Si

Si

RO

(c)

RO

OR
O

OR

O
Si
RO

OR
R'

T3 structure
three dimensional
R' = functional group; R = alkyl group
Fig. 3. Schematic presentation of Ti silane structures [72].

3. Interaction mechanisms between silanes and natural bers


3.1. Fiber surface coating and cell wall modication
In the present literature, there are several different means reported to apply silanes to natural bers. These can be divided into
ber surface treatment and cell wall modication. Spraying is a relatively easy way to treat the ber surface with a silane solution.
The silanes are dissolved into certain organic solvents or solvent/
water mixtures and the prepared solution is directly sprayed onto
the bers. If the solution is water-free, the sprayed silanes can be
partly hydrolyzed via the reaction with water from bers and air.
Due to the fast evaporation of solvent in air, the nano-pores of cell
walls cannot be opened up and thereby the silane molecules will
hardly penetrate into ber cell walls, because it will take many
hours for the chemical to diffuse into dense ber cell walls from
the solution [29]. As a result, spraying only results in a surface
coating with silanes and the inside of cell walls is left untreated.
In one method, silane solution and initiator were directly
pumped into an extruder during an extrusion process of natural bers and thermoplastic matrices. The extruded composites were
subsequently exposed to an environment with high humidity and
temperature (ca. 100% relative humidity and 90 C) in order to
complete the hydrolysis and condensation processes of the silane
[50]. This technique is simple at the initial stage but it will take a
long time at the hydrolysis and condensation stage. In addition,
the silanes added will not only distribute on the interface of ber
and matrices, but some also in the matrices. This causes inefcient
utilization of the silanes added. This process technique is also to be
seen as a surface treatment.
In another approach, thermoplastic matrices were grafted with
silanes and the grafted pre-polymers were used as a new coupling
agent directly blending with natural bers and thermoplastic

810

Y. Xie et al. / Composites: Part A 41 (2010) 806819

matrices [52]. Through this method of application, the alkoxysilanes do not undergo any hydrolysis.
Impregnation processes where natural bers were treated with
prehydrolyzed silane solution to allow the silane penetrate into the
ber lumina and further diffuse into the cell walls have been reported [85,86]. Consequently, both ber surfaces and cell walls
are modied with silanes (bulking treatment). The penetration of
silanes into cell walls is inuenced by the molecular size of silane
which is inuenced by the aging of the hydrolyzed silane solution.
Improper hydrolysis processes may result in a fast condensation of
silanols thereby prematurely increasing the molecular size of the
silanes. In this case diffusion of silanes into the cell walls will be
limited or prevented entirely. The bulking treatment of ber cell
walls can change the properties of cell walls, and as a result promote the performance improvement of the ensuing composites.
Compared to the surface treatment such as spraying, the impregnation process may cause problems for some type of bers. For
example, ne short bers may aggregate and cannot thus evenly
disperse in the solution; the drying process may also consume
energy.
3.2. Interaction mechanisms between silanes and natural bers
3.2.1. Adsorption between silanols and hydroxyl groups of bers
When hydrolyzed silane solutions are mixed with natural bers,
the reactive silanol groups have a high afnity for each other, forming ASiAOASiA bonds and also for the hydroxyl sites of bers via
hydrogen bonds. The adsorption isotherms of silanol groups onto
cellulose ber surfaces have been established using the techniques
of FTIR and UV spectroscopy [87]. The silanols of APS, TAS, or MPS
rstly form a monolayer on the ber, and then are further adsorbed
due to the temperature-driven condensation reaction, which results in the formation of a rigid polysiloxane layer on the ber surface. The polysiloxane layer on the ber surfaces may hinder any
diffusion of the silane molecules into the cell walls, or result in a
concentration gradient in the cell walls. Consequently, the hydrolysis conditions such as pH of solution have to be adjusted to slow
down the condensation thereby allowing the reactive silanol access to hydroxyl sites in the ber cell walls. Both, APS and TAS, displayed higher adsorption towards the surfaces of cellulose bers
than did MPS, since the former two may establish strong hydrogen
bonds between the amino groups and the ber hydroxyl groups.
Owing to steric hindrance by the aromatic ring, c-phenyl-aminopropyltrimethoxy silane (PAPS) shows the lowest adsorption afnity for cellulose bers compared to APS, TAS, and MPS. The
adsorption of silanols in the ber cell walls is not clear in the case
of bulking treatments (impregnation treatment) due to the difculties of in situ determination; however, the afnity of silanols for
the microbrils of the cell walls may be similar to that displayed
with the ber surface. To fully modify the cell walls, the silanol
monomers or oligomers have to be small enough to get access to
the hydroxyl groups of the interior of the cell walls.
3.2.2. Bonding of silanols and the hydroxyl groups of bers
The free silanol groups may undergo further condensation
forming ASiAOASiA network linkages on/in the bers during solvent evaporation as evidenced by 29Si NMR study [72]. Unlike the
SiAOAC bonds, the ASiAOASiA bonds formed are very stable towards hydrolysis [43]. The hydrogen bonds formed between the
adsorbed silanols and hydroxyl groups of natural bers at the
adsorption sites do not, however, convert into the covalent bonds
of ASiAOACA linkages at room temperature [8890], and thus
the adsorbed silanes may be leached off the bers by Soxhlet
extraction with ethanol [87]. Heating (e.g. 110 C for 2 h) may remove the water or solvents in the bers and drive the dehydration
reaction at the adsorption sites between silanols and ber hydroxyl

groups thereby forming ASiAOACA bonds [69,87,91]. Further evidence on the interfacial bonds of silane-modied cellulose bers
measured using X-ray photoelectron spectroscopy (XPS) and FTIR
spectroscopy conrm the occurrence of ASiAOACA bonds
[87,9294]. Heating also promotes the condensation of free silanol
groups resultantly forming the solid polysiloxane layers on the ber surface and the entangled polysiloxane networks in the cell
walls. Hydrogen bonding is also possible between the ASiAOASiA
backbone and the hydroxyl groups of bers. In the case of a bulking
treatment, the polysiloxane networks in the cell walls may reduce
the size of the cell wall nano-pores where the water is normally
able to gain access. As a result, the water sorption of the cell walls
may be reduced. In addition, the aminosilanes such as APS may
form the strong intra- and inter-molecular hydrogen bonds with
its amino group also possessing a strong afnity towards the hydroxyl groups of bers. These interactions in the aminosilane-treated bers may result in the formation of a cage-like structure in the
bers [44,95]. However, it must be emphasized that the bonds of
ASiAOACA are not stable towards hydrolysis under a moist environment. Hydrolysis of ASiAOACA bonds will produce the free silanol groups again. The reproduced silanols may condense
themselves or form ASiAOACA bonds again under heating
conditions.
In general, interaction of silane coupling agents with natural bers may mainly proceed through following steps [72,96]:
(1) Hydrolysis (Fig. 4a): The silane monomers are hydrolyzed in
the presence of water and catalyst (normally acid or base)
liberating alcohol and yielding reactive silanol groups.
(2) Self-condensation (Fig. 4b): During the hydrolysis process,
the concomitant condensation of silanols (aging) also takes
place. The condensation should be minimized at this stage
to leave the silanols free for being adsorbed to the hydroxyl
groups in the natural bers. For the bulking treatment of
bers, the condensation should also be controlled in order
to retain a small molecular size of monomers or oligomers
to diffuse into the cell walls. The condensation rate of silanols is controllable by adjusting the pH of the hydrolysis system. An acidic pH environment is usually preferable to
accelerate the hydrolysis rate of silanes but slow down the
condensation rate of silanols.
(3) Adsorption (Fig. 4c): The reactive silanol monomers or oligomers are physically adsorbed to hydroxyl groups of natural
bers by hydrogen bonds on the ber surfaces (surface coating) and/or in the cell walls (cell wall bulking), which
depends on the molecular size of silanol monomers/oligomers formed. The free silanols also adsorb and react with
each other thereby forming a rigid polysiloxane structures
linked with a stable ASiAOASiA bond.
(4) Grafting (Fig. 4d): Under heating conditions, the hydrogen
bonds between the silanols and the hydroxyl groups of bers
can be converted into the covalent ASiAOACA bonds and
liberating water. The residual silanol groups in the bers will
further condense with each other. The bonds of ASiAOACA
may not be stable towards hydrolysis; however, this bond is
reversible when the water is removed at a raised
temperature.
3.3. Effect of silane treatments on the ber properties
3.3.1. Stabilization against hydrolysis
Although the SiAOAC bond is not stable towards hydrolysis
[43], several authors have found that the silanes in the treated bers or wood are nonetheless resistant against water leaching
[85,87,97]. Commercial microcrystalline cellulosic bers treated
with MPS, APS, and TAS silanes only had slight weight loss after

811

Y. Xie et al. / Composites: Part A 41 (2010) 806819


OR

(a) Hydrolysis:

R'

Si

OH
H or OH

+ 3 H2O

OR

R'

Si

OR

R'

Si

OH

OH

OH

(b) Self-condensation:

+ 3 ROH

OH

OH

+ HO Si

OH

R' Si

R'

OH

OH

OH

O
R'

Si

R'

+ H2O

OH

Si

Si

O
R'

OH
O

R'

+ x H2O

Si

Si

R'

O
OH
OH
O

HO Si

R'

HO

Si

R'

H
OH

(c) Adsorption:

OH

OH
HO

Si

OH
O

Si

OH
R'

HO

Si

OH
O

Si

R'

H
R'

OH

R'

OH

OH
HO

Si

R'

OH

(d) Chemically grafting:

Si

R'

O
OH
O

HO

Si

OH
O

Si

H
R'

OH

R'

Si

R'

Si

+ x H2O
R'

Fig. 4. Interaction of silane with natural bers by hydrolysis process.

15 h Soxhlet extraction with ethanol [87]. Donath et al. [97] determined the anti-leaching ability of wood treated with the oligomers
of methyl- and propyl-triethoxysilane and a multifunctional aqueous siloxane HS 2909 (with the functionalities of alkyl and amino
groups). Both silane oligomers, especially the latter, cannot be leached out of pine wood after treatment. Both silanes MPS and vinyltrimethoxy silane (VTS) can penetrate into and react with the cell
walls of Corsican pine (Pinus nigra) sapwood via radical grafting or
condense in the cell walls [85]. After ve leaching cycles involving
water saturation followed by oven drying, only small amounts of
silane were leached out and the treated wood still maintained a
high anti-swelling efciency (ASE), indicating the silanes still bond
to cell walls after leaching. The high efciency in resisting water
leaching may be apparently attributed to the highly hydrophobic
alkyl groups and/or high afnity of silanols for cellulosic hydroxyl
groups. In case of aminosilane, the strong afnity of amino groups
for the hydroxyl groups of bers should be also considered as a reason. In addition, X-ray mapping (SEM-EDX) of silane treated wood
reveals an accumulation of silicon in the cell lumina and bordered
pits thereby plugging of these typical penetration pathways for
water [97]. Where polymerization occurs within the cell wall, the
silane polymers become entangled with the cell wall polymeric
network and are leach-resistant.
3.3.2. Hydrophobation of bers
By a proper surface and/or bulking treatment with silanes, the
normally hygroscopic natural bers can be converted into a hydrophobic reinforcement for non-polar polymer matrices. A surface
coating only decreases the water sorption rate and does not sub-

stantially reduce the amount of water to be absorbed because


the cell walls are not lled with polymeric silane. In contrast, a bulking treatment may reduce the cell wall nano-pore size and deactivate or mask the hydroxyl functionalities thereby decreasing
water sorption [86]. Hydrophobation treatments can weaken the
hydrogen bonds between natural bers and the agglomeration of
bers is thus reduced during compounding with non-polar polymer matrices.
Treatments of a commercial microcrystalline cellulose ber
with a silane solution of APS, MPS, hexadecyltrimethoxy silane
(HDS), or c-mercaptopropyltrimethoxy silane (MRPS) in an ethanol/water (80/20 v/v) mixture caused a signicant increase in the
water contact angle of treated bers from an original value of
20 (untreated control) up to 110 determined by a dynamic contact angle test [65,66]. The polar component of the ber surface energy is reduced and the non-polar component increases due to the
treatments with alkylsilane such as propyltrimethoxysilane and
HDS [59,66,98100]. Optical microscopy showed that treatment
of ax bers with VTS reduces the heterogeneous nucleation of
PP crystals on ax bers and the treated ber surfaces were
smoother than the untreated bers [99]. Treatment of sisal bers
with 1% MPS in benzene has been reported to reduce the equilibrium moisture content (EMC) of bers from 12.8% to 1.7% and
the water absorption from 123.8% to 64.4% [20]. However, since
benzene does not swell the ber cell wall the penetration of MPS
in the cell walls may consequently be limited and the reduction
in moisture/water sorption may mainly be due to the hydrophobic
effect of the surface coating. Compared to the silanes bearing nonpolar alkyl groups, treatments with APS or PAPS which bear polar

812

Y. Xie et al. / Composites: Part A 41 (2010) 806819

amino groups only resulted in a minor reduction of the moisture


content of jute bers [59]. This may be due to the highly hydrophilic characteristics of the amino end group.
3.3.3. Effect on the thermostabilization
Silane treatment of bers could affect the thermal stability of
natural bers. During the processing of composites, the natural bers are subject to intense heat in the compounding machine or
extruder. Therefore, a positive effect on the thermal properties is
desired. The thermal stability of silane-treated ax bers was studied using thermogravimetric analysis (TGA) at a temperature range
of 30800 C in a helium atmosphere [101]. The bers were treated
by immersing them into the VTS aqueous solution for 1 h to obtain
a weight percent gain of 2.5%. The treated bers exhibited an improved thermal stability of hemicellulose and pectin. Kinetic analysis of the degradation process using Kissinger equation showed an
activation energy which proved to be 22% higher for VTS-treated
bers than the untreated controls.
3.3.4. Effect on the ber strength
The treating system should not contain any ber-damaging element such as an acidic catalyst (for hydrolysis) or include high
temperature. The acidic conditions (ca. pH 4.0), in some cases, for
the acid-catalyzed hydrolysis of silanes were moderate and the
pH was close to that of many natural bers. As a result, silane treatment caused little effect on the ber tensile strength [47,59,102].
Varma et al. [103] reported, however, a 35% and 19% tenacity loss
of coir bers after reuxing in a benzene solution of 0.5% dichloro
methyl vinyl silane (DMVS) or of 2% MPS, respectively (Table 2);
however, these tenacity losses are mostly thought to have been
caused by the benzene (24%), rather than the silane itself (Table 2).
4. Interaction with polymer matrices
4.1. Coupling with thermoplastic matrices
4.1.1. Inter-phase compatibility
The interaction mode between the silane-treated ber and the
polymer matrix is a crucial factor for the mechanical properties
of the resulting NFPCs. Physical blending of the silane-treated bers and the thermoplastic matrices enhances their mutual adherence via inter-molecular entanglement, or acidbase interactions
(ABI) [43]. The interfacial shear strength (ISS) between jute bers
and PP, determined by a microdroplet micromechanical test, was
improved by treating jute bers with a 0.5% APS aqueous solution
[104]. Treatment of aspen chemithermomechanical pulp (CTMP,
mesh 60 lm) with a 4 wt.% ethanol solution of silane VTS or MPS
did not substantially enhance the tensile properties of the resulting
polyethylene composites compared to the untreated composites
[105]. The composites made from the bers treated with silane,
such as HDS or dichlorodiethylsilane (DCS) merely bearing a nonreactive aliphatic chain, displayed only a moderate improvement
in the mechanical properties of the composite [47,53,106,107].

Table 2
Effect of silane treatment on mechanical properties of coir bers [103].
Treatment

Time
(h)

Tenacity (g/
denier)

Elongation at
break (%)

Initial modulus (g/


denier)

Untreated
2% MPS
0.5%
DMVS
0.5%
DMVS
Benzene

1.00
0.50

2.26
1.82
1.47

28.80
23.52
22.44

38.30
36.09
30.13

0.25

1.80

26.12

34.20

1.00

1.71

23.72

35.94

The limited coupling effects on the mechanical strength of the


NFPCs due to the absence of covalent bonds between the silane
functionality and thermoplastic matrices are exemplied in Table 3. The marginal improvements in the mechanical properties
of the composites were mainly attributed to the increased compatibility resulting from the uniform dispersion of silane-treated bers into the thermoplastic matrices [47]. The thermoplastic
molecular chains may also diffuse into the rigid polysiloxane structures on the ber surface forming an entangled interpenetrating
polymer network (IPN) [43,94,108,109].
In the case of aminosilanes, the amino groups cannot react with
the hydrocarbon backbone of PP or PE, but the natural bers and
thermoplastic composites coupled with APS was reported to provide somewhat better mechanical properties than the uncoupled
ones [47,110]. Such improvements have been proposed to be due
to the strong afnity of the amino group towards the hydroxyl
groups of bers and to the formation of a cage-like IPN composed
of the polysiloxane structures [43,44]. The IPN network can entrap
the thermoplastic molecules thereby anchoring the thermoplastic
matrices to the treated ber surface. In addition, acidbase interactions (ABI) may also play a role in determining the interfacial adhesion of composites composed of the aminosilane-treated bers and
specic thermoplastic matrices bearing acidic or basic characteristics, e.g. polystyrene (PS) and polyvinyl chloride (PVC) [47,111].
The calculation according to electron donor and acceptor numbers
of organic molecules indicates that the cellulosic bers treated
with aminosilane have both the acidic (KA) and basic (KB) characteristics, and the basicity is more prominent than the acidity
(KA = 0.33, KB = 0.52) [111]. This bipolar characteristic enhances
the interactions of aminosilane-treated bers with the bipolar
polystyrene (KA = 0.28, KB = 0.46) [111] or PVC (KA = 1.43,
KB = 0.65) [47] since the acidic sites from the aminosilane-treated
bers can interact with the basic sites of thermoplastic matrices
and vice versa. Reinforced polystyrene composites with the cellulose bers treated with different structures of silanes show a linear
increase in the maximum interfacial shear stress (ISS) with the
acidbase interaction value (Iab) between the treated cellulose bers and polystyrene matrix [111]. For the studied silanes with a
functionality of phenyl (C6H5), phenylamino ((CH2)3NHC6H5), amino ((CH2)3NH2) or octadecyl ((CH2)17CH3), the aminosilane-treated
bers display a stronger acidbase interaction with polystyrene
than the other silanes, and hence obtain stronger interfacial adhesion [111]. Similar to a polystyrene matrix, interaction of the acidic
PVC and basic APS also showed acidbase characteristics (Fig. 5)
[47,57]. Compared to the untreated ber/PVC composites, the
PVC composites reinforced with the APS-modied bers displayed
an increase up to 36% in the tensile strength due to the high Iab value between treated ber and PVC; however, treatment with DCS
did not change the Iab of two components of the composite and
thus the tensile strength of resulting PVC composites did not
change [47].
4.1.2. Radical grafting
As presented above, physically blending of the silane-modied
natural bers and thermoplastic matrices only produces a limited
improvement in the mechanical properties of the resulting composites (Table 3). To substantially improve the mechanical properties of natural ber and thermoplastic composites, the
formation of a covalent bond between the silane and non-polar
thermoplastic matrix seems to be necessary [113]. It should, however, be noted that a strong coupling between a reinforcing ber
and the matrix can result in a brittle composite especially with
thermoset matrices [114]. However, most thermoplastic resins
do not bear any reactive functionality. As a result, the silane-treated natural bers cannot react with the target thermoplastic to
form any covalent bond during the compounding process, either

813

Y. Xie et al. / Composites: Part A 41 (2010) 806819


Table 3
Improvement (%) in tensile properties of natural ber/thermoplastic composites coupled with different functionalities of silanes in the absence of initiators.
Silane (%)

VTS (4)
MPS (4)
APS (4)
MPS (2)
HDS (3)
MRPS (3)
APS (0.01)
DCS (0.1)
a

Fiber treatment with silane

Fiber (%)

Immersion in ethanol/water
Immersion in ethanol/water
Immersion in p-xylene
Spray with aqueous solution
Immersion in ethanol/water
Immersion in ethanol/water
Immersion in acetone
Dry blending

Polymer

CTMP (40)
CTMP (40)
CTMP (30)
Wheat straw ber
Cellulosic bers
Cellulosic bers
Pine TMP (10)a
Newsprint ber (45)

LDPE
LDPE
PS
PP
LDPE
LDPE
PE
PVC

Improvement (%)

References

Tensile
strength

Tensile
modulus

Elongation at break

Energy at yield

27
9
6
2
6
12
1
5

65
74
25
10
7
6

15

107
245
18
6

30

16
25
19

[105]
[105]
[44]
[112]
[53]
[53]
[47]
[47]

TMP: thermomechanical pulp.

O
Si (CH2) 3
O

NH2

Cl

O
Si (CH2)3
O

Cl
NH

PVC

APS grafted fiber

Fig. 5. Acidbase reaction between APS-modied bers and PVC matrix [47].

in the mixers or in the extruders. An effective means to covalently


bond the silane to a thermoplastic matrix is by free radical grafting [113,115] or by plasma discharge [116]. The mostly reported
applications are to graft VTS or MPS onto a thermoplastic matrix
in the presence of peroxide initiators such as benzoyl peroxide
(BPO) or dicumyl peroxide (DCUP) [e.g. 44,49]. At an elevated
temperature the peroxide rst decomposes, generating oxy radicals. The oxy radicals not only have the potential to abstract
hydrogen from the backbone of thermoplastic molecules or natural bers, but can also add to vinyl double bonds of vinylsilane,
producing vinyl radicals. The vinyl free electron may either combine with each other (homo-polymerization) or attack the other

molecules in a similar fashion to propagate the free radical reaction. A previous study showed, however, that VTS did not homopolymerize during reaction due to its bulky silane side groups
[117]. Accordingly, the radical reaction would ultimately result
in a grafting of vinylsilane onto thermoplastics as shown in
Fig. 6 [44,118]. The grafting of vinylsilanes onto thermoplastic
matrices has two options: one approach is to graft vinylsilanes
onto matrices with the resulting copolymer being used as a coupling agent to bond the natural ber and matrices; the other way
is to treat the ber with the vinylsilane solution and the grafting
reaction appears in the thermal compounding process of vinylsilane treated bers and thermoplastic matrices.

CH 2

CH 2
R-O-O-R

CH 2

CH 2

HC

CH 2

PE chain

PE radical

OR
OC 2 H 5
R-O-O-R
H2C

CH

Si

OC 2 H 5

CH 2

. CH

OC 2 H 5

HC

CH 2

CH 2
+

. CH

Si

OC 2 H 5

OC 2 H 5
OR

OR
CH 2

OC 2 H 5

CH 2 CH 2

OC 2 H 5

OC 2 H 5

Grafting
Si

OC 2 H 5

OC 2 H 5

HC
CH 2

CH

Si

OC 2 H 5

OC 2 H 5

Fig. 6. Radical grafting of vinylsilane onto polyethylene matrix.

814

Y. Xie et al. / Composites: Part A 41 (2010) 806819

Table 4
Mechanical properties of natural ber/thermoplastic composites coupled with the representative vinylsilanes in the presence of peroxide initiator.
Components of composite
Fiber (%)

Polymer

Silane (%)

Initiator (%)

Henequen (20)
CTMP of aspen (30)
Wood our (40)

HDPE
LDPE
HDPE

VTS (1)
MPS (4)
VTS (2)

DCUP (0.5)
DCUP (unclear)
DCUP (0.17)

Flax (30)
CTMP of aspen (30)
CTMP of spruce and balsam r (20)

PP
LDPE
PVC

MPS (3)
VTS (4)
MRPS (2)

DCUP (0.3)
DCUP (unclear)
BPO (0.8)

4.1.2.1. Copolymer as new coupling agent. Grafting of VTS alone onto


low density polyethylene (LDPE) with different concentrations of
VTS (027.33 part per hundred (phr)) and BPO (01.25 phr) has revealed that the VTS grafted polyethylene produced a maximum
tensile strength at the concentrations of 5.0 phr VTS and 0.25 phr
BPO [118]. The grafting efcacy of VTS and maleic anhydride onto
PP in the presence of DCUP were studied and the results demonstrated that VTS has a higher reactivity towards PP than maleic
anhydride [61,119]. When the copolymers of VTS grafted polypropylene (VSPP) and MAPP were respectively used as a new coupling
agent to couple wood our and PP matrix, VSPP coupled composites were 47% stronger in tensile strength and 35% lower in water
uptake than the composites coupled with the copolymer MAPP
[52]. The study from Arbelaiz et al. [120] also reveals a superiority
of VSPP to MAPP used as the coupling agents. MAPP has been
extensively studied and is well known to be an excellent coupling
agent in the bonding of natural bers and thermoplastic matrices.
This provides a means to further improve the interfacial properties
of the NFPCs by using the VSPP copolymer.

References

ca. 30% in tensile strength, 50% in Iosipescu shear strength


38% in tensile strength, 13% in tensile modulus
87% in exural strength, 18.8% in exural modulus,
more than 100% in impact strength
More than 60% in tensile strength
66% in tensile strength, 44% in tensile modulus
98% in tensile strength

[91,94,124]
[51]
[50,121,122]

OR
RO

Si

[106]
[51]
[57]

O
R'

OR

N3

Fig. 7. A general chemical formula of azidosilane. Representatively, R0 = arylcontaining group.

tional azido group can react with thermoplastic matrices forming


covalent bonds (Fig. 8). Upon heating, the azide may decompose
forming a highly reactive intermediate nitrene and simultaneously
releasing a nitrogen molecule. The nitrene can react with any of the
CAH bonds in the thermoplastic polymer by inserting itself into
the carbonhydrogen single bond or by adding to a carboncarbon
double bond or aromatic system [126]. The required temperature
for the covalent reaction is below 200 C (in the temperature range
of thermal process of NFPCs) and the reaction does not need any
peroxide initiator. Accordingly, azidosilanes would also be of great
interest to couple natural bers and polymer matrices. However,
the application of azidosilanes on the NFPCs is little reported.
4.2. Coupling with thermosetting resins
In contrast with thermoplastic resins, the thermosetting resins
(thermosets) are usually liquids prior to curing, bear reactive
organofunctionalities and thereby are able to form three-dimensional networks after a curing reaction. Compared to thermoplastics, thermoset resins are more reactive towards the
organofunctionalities of silanes. Unlike the extrusion process of
natural ber/thermoplastic composites where there is a shear failure of long bers, the long natural bers can be easily incorporated
into the thermosetting resins without any damage from the pro-

OR

OH

RO Si

OR
R"

N3

OR

OR
O

Si

N3

Si

R"

N:

OR

OR

R"

N3

OR
R"

OR

Si
OR

Si

R"

N:

N2

R"

NH

OR

OR

CH2
O
CH2
n

Si
OR

CH
CH2
n

4.1.2.2. Grafting reaction between the silane-treated ber and matrices. A study of the melt rheological behavior showed that treatment of pineapple bers with VTS (4.0 wt.% of ber) and
peroxide DCUP (2.0 wt.% of ber) caused an increase in the melt
viscosity of the resulting LDPE composites. This was attributed to
the increased bermatrix interfacial interaction due to a grafting
reaction. Treatment of pineapple bers with the peroxides DCUP
(0.5 wt.% of LDPE) or BOP (1.0 wt.% of LDPE) alone also increased
the melt viscosity of resulting LDPE composites due to the free radical grafting reaction between bers and LDPE at a high temperature (125145 C) [49]. The grafting reaction of VTS-treated wood
our with PE in the presence of DCUP initiator caused a signicant
increase in the exural strength (up to 87%) and impact strength
(more than 100%) of the resulting extruded composites
[50,121,122]. The reported effects of grafting vinylsilanes with different thermoplastic matrices on the mechanical properties of
resulting NFPCs are exemplied in Table 4.
Grafting effects of vinylsilane on the mechanical properties of
the resulting NFPCs depended on the types of thermoplastic matrices [123,124]. Treatment of bers with VTS or MPS in the presence
of DCUP initiator resulted in an increase of up to 62% in the tensile
strength of the PE matrix composites [51]; however, incorporation
of bers with the same treatment into a polystyrene matrix did not
substantially improve the mechanical properties of the ensuing
composites compared to the untreated ber reinforced composites
[44]. It may be concluded that the proper match of silane organofunctionality with the target polymer matrices is important to attain a desired coupling effect (Table 1).
Azidosilanes with a general formula (CH3O)3SiAR0 ASO2N3
(Fig. 7) have been reported to be highly efcient in coupling mineral llers with polyolen matrices [6264,125]. Azidosilanes can
also be hydrolyzed thereby forming the reactive silanols which
are able to react with or condense in the bers. The organofunc-

Improvement in mechanical properties (%)

Fig. 8. A proposed coupling reaction between azidosilane-treated bers and


thermoplastic matrices [63].

815

Y. Xie et al. / Composites: Part A 41 (2010) 806819

OH
O

n
Fig. 9. Molecular structure of epoxy pre-polymer diglycidyl ethers of bisphenol A
(DGEBA).

cessing. In the published literature, the most reported thermosets


for NFPCs are epoxy compounds and unsaturated polyesters
[1,114,127130].
4.2.1. Reaction with epoxy resin
Bleached soda pulp bers were modied with several silanes
(3% of APS, MPS, HDS, or MRPS) and then used as a reinforcement
(40 vol.%) for epoxy resin. The results show that APS was the most
efcient in improving the mechanical properties and reducing the
water absorption of the resulting composites [45]. The reaction
kinetics of epoxy pre-polymer diglycidyl ethers of bisphenol A
(DGEBA, as shown in Fig. 9) with APS was systematically studied
using differential scanning calorimetry (DSC), FTIR, and size exclusion chromatography (SEC), indicating that the epoxy is highly
reactive towards the amine group producing an insoluble fraction
after reaction [131]. The insoluble fraction is formed via the gelation of silane ethoxy and methylol groups which are produced by

the amine-epoxy reaction [46]. The interfacial interactions of DGEBA and silanes were studied by incorporating the polysiloxane
powders formed from condensation of hydrolyzed APS into DGEBA
matrix [108,132]. FTIR evidence showed that the epoxy resin reacted with polyaminosiloxane forming covalent bonds at the interface [132]. Use of a polymerizing accelerator benzyl
dimethylamine (BDMA) increased the reaction rate of the epoxy
resin and polyaminosiloxane. The coupling mechanism of aminosilanes in the ber/epoxy composites have been proposed by several authors as summarized in Fig. 10 [e.g. 45,46,108,132].
Treatments of ax bers with APS resulted in an enhancement of
tensile strength up to 17%, and of the modulus up to 25%, but did
not change the impact toughness of the resulting epoxy composites [133]. Sisal ber/epoxy composites (40 vol.% of ber) coupled
with 5% APS showed a considerable improvement in the compression strength and wet tensile strength [19].
4.2.2. Reaction with polyester resin
The vinylsilanes have been demonstrated to be able to efciently couple natural bers and unsaturated polyester (UP) resins.
A dominant co-polymerization mechanism is proposed where the
methacrylic or vinyl group in the silanes is capable of reacting with
the double bonds of the UP matrix when the mixture system is initiated by a peroxide initiator (Fig. 11). Bleached soda pulp bers
treated with 3% silane of APS or MPS (ber content: 40 vol.%) were
incorporated into an unsaturated glycerophthalic polyester
OH

H
O
O Si (CH2)3 N
O
H

CH2
O
O Si (CH2)3 N
O
CH2

H
C

2 H2C
O

CH
CH
OH

epoxy resin

APS grafted fiber

Fig. 10. Coupling reaction between APS-grafted natural ber with epoxy resin.

Si

O
(CH2)3 O C C CH2
O
CH3

HC

+ m

UP resin

O C CH

CH2

CH

CH2

C O

CH2
O

CH3

O C CH CH

O
Si (CH2)3
O

Styrene

MPS grafted fiber

O C CH CH C O

CH2
O

peroxide

Fig. 11. Reaction of MPS-grafted ber with unsaturated polyester resin in the presence of a peroxide initiator.

816

Y. Xie et al. / Composites: Part A 41 (2010) 806819

(60 wt.% in styrene) with benzoyl peroxide used as a free radical


initiator [45]. Treatment with APS caused a slight increase in the
tensile Youngs modulus of the composite, but did not change the
other mechanical properties. The results are consistent with a previous study [134]. This is due to the lack of reactivity between the
amino groups of APS and the polyester matrix. In contrast to APS,
treatment of bers with MPS improved the exural strength of
the resulting composites and slightly reduced the water uptake
in a 4-week water immersion test [45].
Compared to bleached soda pulp bers [45], the application of
MPS (15%) to the other types of natural bers such as que and
sisal bers resulted in an enhancement in the mechanical properties of the resulting UP matrix composites. In the presence of the
radical initiator methyl ethyl ketone peroxide (MEKP), the tensile
strength and modulus of the que ber/UP composite increase
up to 60% and 80%, respectively, due to the MPS treatment of bers
[135]. The exural strength of sisal/UP composites can be improved up to 63% [20]. Comparably, the tensile strength of pineapple leaf ber (PALF)/UP composites coupled with VTS increased
40% as compared to the uncoupled composites [134]. Dynamic
mechanical analysis (DMA) showed that treatment of banana bers with MPS or VTS caused an increase in the dynamic storage
modulus and a reduction in the damping value of the resulting
polyester composites, suggesting the interfacial adhesion between
banana ber and the polyester is improved [54,136]. The varying
mechanical improvement due to the different ber types after
treatment with MPS may be due to the different ber constituents
and preparation methods of composites in the studies.

5. Outdoor performance
The moisture sorption of NFPCs measured according to standard
ASTM D-1037 [137] generally indicates very low moisture contents
(in a range of ca. 2%), although this depends upon the ber contents. This has been attributed to the effect of the encapsulation
of bers by polymers [138]; however, with the NFPCs in service
there often appears deleterious effects such as surface colour fading and erosion, warpage, mold growth, fungal decay, and strength
loss after a long-term exposure to an external environment
[21,139144]. These issues are mostly associated with the moisture sorption of the natural bers. The moisture sorption of NFPCs
in service is slow and seldom reaches an equilibrium condition in a
moist environment. The core of the composites may have very low
moisture contents whilst the surface layers may be signicantly
saturated with water [145]. However, even though it may take
time for the water to penetrate the core of such composites, the
water susceptibility of natural bers remains a cause of concern,
which is particularly important if the bers are exposed to moisture either through damage or cutting of the composite. Consequently, hydrophobation treatments of bers thereby reducing
the moisture sorption are of importance for the NFPCs during environmental exposure in order to reduce potential damage associated with water.
Oil palm bers modied with VTS in an alcoholwax mixture
(60:40) were soaked in water at different temperatures (30, 50,
70, and 90 C) to determine their water uptake behavior and the results showed that VTS-treatment reduced the water uptake of bers at all temperatures used [146]. This makes the resulting
polymer composites drier under a moist environment thereby
reducing the risk of environmental damage such as deformation
and fungal decay [147]. Epoxy composites reinforced with sisal
which had been treated with APS or with jute ber treated with
c-glycidoxypropyltrimethoxy silane (GPS) exhibited a reduced susceptibility of moisture on the exural strength [19,58]. Coir or oil
palm bers modied with MPS were incorporated into polyester

resin and the modifying effects on the resulting composites properties of water resistance and fungal decay were respectively
examined during one year weathering [148]. The mass loss due
to fungal decay reached over 15% for unmodied composites. In
comparison, MPS modied composites only lost 6% of the mass.
The modied composites after fungal test obtained a moisture content of 6%, less than the 12% observed with the unmodied composites. The low moisture content may be related to the reduced
mass loss due to fungal decay [147]. The tensile and exural
strength of the modied composites were slightly reduced (up to
8%), which is distinctly less than over 30% for the unmodied
controls.
6. Summary
Most established silanes used for natural ber/polymer composites are trialkoxysilanes bearing a non-reactive alkyl or reactive
organofunctionality. Silane is hydrolyzed forming reactive silanols
and is then adsorbed and condensed on the ber surface (solgel
process) at a specic pH and temperature. The hydrogen bonds
formed between the adsorbed silanols and hydroxyl groups of natural bers may be further converted into covalent bonds by heating the treated bers at a high temperature, although such bonds
are susceptible to hydrolysis.
The interaction modes of the silane and matrix are dominated
by the organofunctionality of silane and the matrix characteristics.
Physical compatibility (such as molecular entanglement, or acid
base interactions) between silane-grafted ber and thermoplastic
matrices only provides a limited improvement in the mechanical
properties of the resulting composites. To substantially improve
the interfacial adhesion, a chemical bonding between the organofunctionalities of silanes and the matrices is required. For the inert thermoplastic matrices, a free radical process is an effective
means to couple the vinylsilane treated ber and matrices. In the
cases of thermoset matrices, the organofunctionalities of silanes
can react with the functional groups of thermoset matrices in the
presence of catalysts or radical initiators. Proper treatment of bers with silane can increase the interfacial adhesion to the target
polymer matrices and improve the mechanical and outdoor performance of the resulting ber/polymer composites.
Acknowledgement
The author Dr. Yanjun Xie would like to thank the German Academic Exchange Service (DAAD) for a research grant support.
References
[1] Hill CAS, Abdul Khalil HPS. Effect of ber treatments on mechanical properties
of coir or oil palm ber reinforced polyester composites. J Appl Polym Sci
2000;78:168597.
[2] Zadorecki P, Michell AJ. Future prospects for wood cellulose as reinforcement
in organic polymer composites. Polym Compos 1989;10:6977.
[3] Bledzki AK, Gassan J. Composites reinforced with cellulose based bres. Prog
Polym Sci 1999;24:22174.
[4] Lu JZ, Wu Q, McNaabb HS. Chemical coupling in wood ber and polymer
composites: a review of coupling agents and treatments. Wood Fiber Sci
2000;32:88104.
[5] Mukhopadhyay S, Deopura BL, Alagiruswamy R. Interface behavior in
polypropylene composites. J Thermoplast Compos 2003;16:47995.
[6] George J, Sreekala MS, Thomas S. A review on interface modication and
characterization of natural ber reinforced plastic composites. Polym Eng Sci
2001;41:147185.
[7] Narkis M, Chen JH. Review of methods for characterization of interfacial
bermatrix interactions. Polym Compos 1988;9:24551.
[8] Ishida H. A review of recent progress in the studies of molecular and
microstructure of coupling agents and their functions in composites, coatings
and adhesive joints. Polym Compos 1984;5:10123.
[9] Jiang H, Kamdem DP. Development of poly(vinyl chloride)/wood composites.
A literature review. J Vinyl Addit Technol 2004;10:5969.

Y. Xie et al. / Composites: Part A 41 (2010) 806819


[10] Li Y, Mai YW, Ye L. Sisal bre and its composites: a review of recent
developments. Compos Sci Technol 2000;60:203755.
[11] Sabu
T,
Pothan
L.
Cellulose
bre
reinforced
polymer
composites. Philadelphie: Old City Publishing; 2009.
[12] Xiao Z, Zhao LB, Xie Y, Wang QW. Review for development of wood plastic
composites. J Northeast Forest Univ 2003;31:8993.
[13] Seymour RB. Cellulose-lled polymer composites. Pop Plast 1978;23:2730.
[14] Wambua P, Ivens J, Verpoest I. Natural bres: can they replace glass in bre
reinforced plastics? Compos Sci Technol 2003;63:125964.
[15] Bledzki AK, Gassan J, Theis S. Wood-lled thermoplastic composites. Mech
Compos Mater 1998;34:5638.
[16] Cantero G, Arbeliaz A, Liano-Ponte R, Mondargon I. Effects of bre treatment
on wettability and mechanical behavior of ax/polypropylene composites.
Compos Sci Technol 2003;63:124754.
[17] Raj RG, Kokta BV. Compounding of cellulose bers with polypropylene: effect
of ber treatment on dispersion in the polymer matrix. J Appl Polym Sci
1989;38:198796.
[18] Kazayawoko M, Balatinecz JJ. Matuana LM surface modication and adhesion
mechanisms in woodberpolypropylene composites. J Mater Sci
1999;34:618999.
[19] Bisanda ETN, Ansell MP. The effect of silane treatment on the mechanical and
physical properties of sisalepoxy composites. Compos Sci Technol
1991;41:16578.
[20] Singh B, Gupta M, Verma A. Inuence of ber surface treatment on the
properties of sisalpolyester composites. Polym Compos 1996;17:9108.
[21] Schirp A, Wolcott M. Inuence of fungal decay and moisture absorption on
mechanical properties of extruded woodplastic composites. Wood Fiber Sci
2005;37:64352.
[22] Li R, Ye L, Mai YW. Application of plasma technologies in bre-reinforced
polymer composites: a review of recent developments. Compos Part A Appl
Sci 1997;28:7386.
[23] Sun D. Investigating the plasma modication of natural ber fabrics-the
effect on fabric surface and mechanical properties. Text Res J 2005;75:
63944.
[24] Belgacem MN, Bataille P, Sapieha S. Effect of corona modication on the
mechanical properties of polypropylene/cellulose composites. J Appl Polym
Sci 1994;53:37985.
[25] Sakata I, Morita M, Tsuruta N, Morita K. Activation of wood surface by corona
treatment to improve adhesive bonding. J Appl Polym Sci 1993;49:12518.
[26] Hill CAS. Woodplastic composites: strategies for compatibilising the phases.
J Inst Wood Sci 2000;15:1406.
[27] Daneault C, Kokta BV, Maldas D. Grafting of vinyl monomers onto wood bers
initiated by peroxidation. Polym Bull 1988;20:13741.
[28] Hong CK, Kim N, Kang SL, Nah C, Lee YS, Cho BH, et al. Mechanical properties
of maleic anhydride treated jute bre/polypropylene composites. Plast
Rubber Compos 2008;37:32530.
[29] Hill CAS. Wood modication: chemical, thermal and other processes. John
Wiley & Sons, Ltd.; 2006.
[30] Belgacem MN, Gandini A. The surface modication of cellulose bres for use
as reinforcing elements in composite materials. Compos Interf
2005;24:4175.
[31] Mohanty AK, Dryal LT, Misra M. Novel hybrid coupling agent as an adhesion
promoter in natural ber reinforced powder polypropylene composites. J
Mater Sci Lett 2002;21:18858.
[32] Sun ZY, Han HS, Dai GC. Mechanical properties of injection-molded natural
ber-reinforced polypropylene composites: formulation and compounding
processes. J Reinf Plast Compos 2009. doi:10.1177/0731684408100264, 2009.
[33] Gassan J, Bledzki AK. The inuence of ber-surface treatment on the
mechanical properties of jutepolypropylene composites. Compos Part A
Appl Sci 1997;28:10015.
[34] Felix JM, Gatenholm P. The nature of adhesion in composites of modied
cellulose bers and polypropylene. J Appl Polym Sci 1991;42:60920.
[35] Botaro VR, Gandini A, Belgacem Naceur. Heterogeneous chemical
modication of cellulose for composite materials. J Thermoplast Compos
2005;18:10717.
[36] Maldas D, Kokta BV. Improving adhesion of wood ber with polystyrene by
chemical treatment of ber with coupling agent and the inuence on the
mechanical properties of the composites. J Adhes Sci Technol 1989;3:52939.
[37] Richelt L, Poller S. Uber die umsetzung von cellulose und lignin mit
isocyanates bzw. Isocyanatgruppenhaltigen prapolymeren. Acta Polym
1981;32:1726.
[38] Rider AN, Arnott DR. Boiling water and silane pre-treatment of aluminium
alloys for durable adhesive bonding. Int J Adhes Adhes 2000;20:20920.
[39] Wu HF, Dwight DW, Huff NT. Effects of silane coupling agents on the
interphase and performance of glassber-reinforced polymer composites.
Compos Sci Technol 1997;57:97583.
[40] Clark HA, Plueddemann EP. Bonding of silane coupling agents in glassreinforced plastics. Mod Plast 1963;40. p. 1335, 1378, 1956.
[41] Park JM, Subramanian RV, Bayoumi AE. Interfacial shear strength and
durability improvement by silanes in single-lament composite specimens
of basalt ber in brittle phenolic and isocyanate resins. J Adhes Sci Technol
1994;8:13350.
[42] Favis BD, Blanchard LP, Leonard J, PrudHomme RE. The interaction of a
cationic silane coupling agent with mica. J Appl Polym Sci 2003;28:123544.
[43] Plueddemann EP. Silane coupling agents. 2nd ed. New York and
London: Plenum Press; 1991.

817

[44] Maldas D, Kokta BV, Daneault C. Inuence of coupling agents and treatments
on the mechanical properties of cellulose berpolystyrene composites. J
Appl Polym Sci 1989;37:75175.
[45] Abdelmouleh M, Bou S, Belgacem MN, Dufresne A, Gandini A. Modication
of cellulose bers with functionalized silanes: effect of the ber treatment on
the mechanical performances of cellulosethermoset composites. J Appl
Polym Sci 2005;98:97484.
[46] Serier A, Pascault JP, Lam TM. Reaction in aminosilaneepoxy prepolymer
systems. II. Reactions of alkoxysilane groups with or without the presence of
water. J Polym Sci Polym Chem 1991;29:112531.
[47] Matuana LM, Woodhams RT, Balatinecz JJ, Park CB. Inuence of interfacial
interactions on the properties of PVC/cellulosic ber composites. Polym
Compos 1998;19:44655.
[48] George J, Bhagawan SS, Thomas S. Thermogravimetric and dynamic
mechanical thermal analysis of pineapple bre reinforced polyethylene
composites. J Therm Anal 1996;47:112140.
[49] George J, Janardhan R, Anand JS, Bhagawan SS, Thomas S. Melt rheological
behaviour of short pineapple bre reinforced low density polyethylene
composites. Polymer 1996;37:542131.
[50] Bengtsson M, Oksman K. Silane crosslinked wood plastic composites:
processing and properties. Compos Sci Technol 2006;66:217786.
[51] Raj RG, Kokta BV, Maldas D, Daneault C. Use of wood bers in thermoplastics.
VII. The effect of coupling agents in polyethylenewood ber composites. J
Appl Polym Sci 1989;37:1089103.
[52] Nachtigall SMB, Cerveira GS, Rosa SML. New polymeric-coupling agent for
polypropylene/woodour composites. Polym Test 2007;26:61928.
[53] Abdelmouleh M, Bou S, Belgacem MN, Dufresne A. Short natural-bre
reinforced polyethylene and natural rubber composites: effect of silane
coupling agents and bres loading. Compos Sci Technol 2007;67:162739.
[54] Pothan LA, Thomas S, Groeninckx G. The role of bre/matrix interactions on
the dynamic mechanical properties of chemically modied banana bre/
polyester composites. Compos Part A Appl Sci 2006;37:12609.
[55] Ismail H. The effect of ller loading and a silane coupling agent on the
dynamic properties and swelling behaviour of bamboo lled natural rubber
compounds. J Elastom Plast 2003;35:14959.
[56] Ismail H, Shuhelmy S, Edyham MR. The effects of a silane coupling agent on
curing characteristics and mechanical properties of bamboo bre lled
natural rubber composites. Eur Polym J 2002;38:3947.
[57] Beshay A, Hoa SV. Reinforcement of polyvinyl chloride (PVC) and polystyrene
(PS) with cellulose bers treated with silane. J Thermoplast Compos
1990;3:26474.
[58] Gassan J, Bledzki AK. Effect of moisture content on the properties of silanized
juteepoxy composites. Polym Compos 1997;18:17984.
[59] Doan TTL. Investigation on jute bres and their composites based on
polypropylene and epoxy matrices. PhD thesis, Technischen Universitt
Dresden; 2006.
[60] Pickering KL, Abdalla A, Ji C, McDonald AG, Franich RA. The effect of silane
coupling agents on radiata pine bre for use in thermoplastic matrix
composites. Compos Part A Appl Sci 2003;34:91526.
[61] Nachtigall SMB, Stedile FC, Felix AHO, Mauler RS. Polypropylene
functionalization with vinyltriethoxysilane. J Appl Polym Sci 1999;72:
13139.
[62] Miller JD, Ishida H. Controlling and monitoring interfacial reactions in
composites of azidosilane modied glass lled polyethylene. Polym Compos
1988;9:129.
[63] Miller JD, Ishida H, Maurer FHJ. Dynamic-mechanical properties of interfacially
modied glass sphere lled polyethylene. Rheol Acta 1988;27:397404.
[64] McFarren GA, Sanderson TF, Schappell FG. Azidosilane polymerller
coupling agent. Polym Eng Sci 1977;17:469.
[65] Gliesche K, Mder E. Langfaserverstrkte Kunststoffe auf der Basis von
Naturfasern. In: Proceedings of 7th international techtextil symposium,
Frankfurt, Germany; 1995.
[66] Abdelmouleh M, Bou S, Belgacem MN, Duarte AP, Ben Salah A, Gandini A.
Modication of cellulosic bres with functionalised silanes: development of
surface properties. Int J Adhes Adhes 2004;24:4354.
[67] Hertl W. Mechanism of gaseous siloxane reaction with silica. I. J Phys Chem
1968;72:124853.
[68] Krasnoslobodtsev AV, Smirnov SN. Effect of water on silanization of silica by
trimethoxysilanes. Langmuir 2002;18:31814.
[69] Castellano M, Gandini A, Fabbri P, Belgacem MN. Modication of cellulose
bres with organosilanes: under what conditions does coupling occur? J
Colloid Interf Sci 2004;273:50511.
[70] Schneider MA, Brebner KI. Woodpolymer combinations: the chemical
modication of wood by alkoxysilane coupling agents. Wood Sci Technol
1985;19:6773.
[71] Matuana LM, Balatinecz JJ, Park CB, Sodhi RNS. X-ray photoelectron
spectroscopy study of silane-treated newsprint-bers. Wood Sci Technol
1999;33:25970.
[72] Salon MCB, Gerbaud G, Abdelmouleh M, Bruzzese C, Bou S, Belgacem MN.
Studies of interactions between silane coupling agents and cellulose bers
with liquid and solid-state NMR. Magnet Reson Chem 2007;45:47383.
[73] Kang HJ, Meesiri W, Blum FD. NMR studies of the hydrolysis and molecular
motion of aminopropylsilane. Mater Sci Eng A Struct 1990;126:26570.
[74] Miller AC, Berg JC. Effect of silane coupling agent adsorbate structure on
adhesion performance with a polymeric matrix. Compos Part A Appl Sci
2003;34:32732.

818

Y. Xie et al. / Composites: Part A 41 (2010) 806819

[75] Wacker Chemical AG publication. Products information brochure:


organofunctional silanes from Wacker. <http://www.wacker.com/>.
[76] Tesoro G, Wu Y. Silane coupling agents: the role of the organofunctional
group. J Adhes Sci Technol 1991;5:77184.
[77] Navoroj S, Culler R, Koenig JL, Ishida H. Structure and adsorption
characteristics of silane coupling agents on silica and E-glass ber;
dependence on pH. J Colloid Interf Sci 1984;97:30917.
[78] Riegel B, Blittersdorf S, Kiefer W, Hofacker S, Mueller M, Schottner G. Kinetic
investigations
of
hydrolysis
and
condensation
of
the
glycidoxypropyltrimethoxysilane/aminopropyltriethoxy-silane system by
means of FT-Raman spectroscopy I. J Non-Cryst Solids 1998;226:7684.
[79] Pohl ER, Osterholtz FD. Kinetics and mechanism of aqueous hydrolysis and
condensation of alkyltrialkoxysilanes. In: Ishida H, editor. Molecular
characterization of composite interfaces. New York: Plenum Press; 1983.
[80] Pantoja M, Daz-Benito B, Velasco F, Abenojar J, del Real JC. Analysis of
hydrolysis process of c-methacryloxypropyltrimethoxysilane and its
inuence on the formation of silane coatings on 6063 aluminum alloy. Appl
Surf Sci 2009;255:638690.
[81] Chiang CH, Ishida H, Koenig J. The structure of c-aminopropyltriethoxysilane
on glass surfaces. J Colloid Interf Sci 1980;74:396404.
[82] Leyden DE, Atwater JB. Hydrolysis and condensation of alkoxysilanes
investigated by internal reection FTIR spectroscopy. J Adhes Sci Technol
1991;5:81529.
[83] Salon MCB, Abdelmouleh M, Bou S, Belgacem MN, Gandini A. Silane
adsorption onto cellulose bers: hydrolysis and condensation reactions. J
Colloid Interf Sci 2005;289:24961.
[84] Salon MCB, Bayle PA, Abdelmouleh M, Bou S, Belgacem MN. Kinetics of
hydrolysis and self-condensation reaction of silanes by NMR spectroscopy.
Colloid Surf A 2008;312:8391.
[85] Hill CAS, Mastery Farahani MR, Hale MDC. The use of organo alkoxysilane
coupling agents for wood preservation. Holzforschung 2004;58:31625.
[86] Donath S, Militz H, Mai C. Creating water-repellent effects on wood by
treatment with silanes. Holzforschung 2006;60:406.
[87] Abdelmouleh M, Bou S, ben Salah A, Belgacem MN, Gandini A. Interaction of
silane coupling agents with cellulose. Langmuir 2002;18:32038.
[88] Daniels MW, Francis LF. Silane adsorption behavior, microstructure, and
properties of glycidoxypropyltrimethoxysilane-modied colloidal silica
coatings. J Colloid Interf Sci 1998;205:191200.
[89] Nishiyama N, Horie K, Asakura T. Adsorption behavior of a silane coupling
agent onto a colloidal silica surface studied by 29Si NMR spectroscopy. J
Colloid Interf Sci 1989;129:1139.
[90] Vrancken KC, Coster LD, Voort PVD, Grobet PJ, Vansant EJ. The role of silanols
in the modication of silica gel with aminosilanes. J Colloid Interf Sci
1995;170:717.
[91] Valadez-Gonzalez A, Cervantes-Uc JM, Olayo R, Herrera-Franco PJ. Effect of
ber surface treatment on the ber-matrix bond strength of natural ber
reinforced composites. Compos Part B Eng 1999;30:30920.
[92] Matias MC, De La Orden MU, Gonzalez-Sanchenz C, Martinez-Urreaga J.
Comparative spectroscopic study of the modication of cellulosic materials
with different coupling agents. J Appl Polym Sci 2000;75:25666.
[93] Valadez-Gonzalez A, Cervantes-Uc JM, Olayo R, Herrera-Franco PJ. Chemical
modication of henequen bers with an organosilane coupling agent.
Compos Part B Eng 1999;30:32131.
[94] Herrera-Franco PJ, Valadez-Gonzalez A. A study of the mechanical properties
of short natural-ber reinforced composites. Compos Part B Eng
2005;36:597608.
[95] Kokta BV, Maldas D, Daneault C, Beland P. Composites of polyvinyl chloridewood bers. III: effect of silane as coupling agent. J Vinyl Technol 1990;12:
14653.
[96] Arkles B, Steinmetz JR, Zazyczny J, Mehta P. Factors contributing to the
stability of alkoxysilanes in aqueous solution. J Adhes Sci Technol 1992;6:
193206.
[97] Donath S, Militz H, Mai C. Wood modication with alkoxysilanes. Wood Sci
Technol 2004;38:55566.
[98] Van de Velde K, Kiekens P. Inuence of ber surface characteristics on the
ax/polypropylene interface. J Thermoplast Compos 2001;14:24460.
[99] Biagiotti J, Puglia D, Torre L, Kenny JM. A systematic investigation on the
inuence of the chemical treatment of natural bers on the properties of their
polymer matrix composites. Polym Compos 2004;25:4709.
[100] Mohammed-Ziegler I, Marosi G, Matko S, Horvlgyi Z, Toth A. Silylation of
wood for potential protection against biodegradation. An ATR-FTIR, ESCA and
contact angle study. Polym Adv Technol 2003;14:7905.
[101] Arbelaiz A, Fernndez B, Ramos JA, Mondragon I. Thermal and crystallization
studies of short ax bre reinforced polypropylene matrix composites: effect
of treatments. Thermochim Acta 2006;440:11121.
[102] Sreekala MS, Kumaran MG, Thomas S. Oil palm bers: morphology, chemical
composition, surface modication, and mechanical properties. J Appl Polym
Sci 1997;66:82135.
[103] Varma DS, Varma M, Varma IK. Coir ber II: evaluation as a reinforcement in
unsaturated polyester resin composites. J Reinf Plast Comp 1985;4:41931.
[104] Park JM, Kim PG, Jang JH, Wang Z, Hwang BS, DeVries KL. Interfacial
evaluation and durability of modied Jute bers/polypropylene (PP)
composites using micromechanical test and acoustic emission. Compos Part
B Eng 2008;39:104261.
[105] Bashay AD, Kokta BV. Use of wood bers in thermoplastic composites II:
polyethylene. Polym Compos 1985;6:26171.

[106] Mieck KP, Nechwatal A, Knobelsdorf C. Faser-matrix-haftung in


kunststoffverbunden aus thermoplastischer matrix und achs. 1. Die
ausrstung mit silanen. Die Angew Makromol Chem 1995;224:7388.
[107] Xie Y. Silane treated wood our as reinforcement for thermoplastic
composites. Project report of German academic exchange service, GerogAugust-University Goettingen, Germany; 2008.
[108] Jensen RE, Palmese GR, Mcknight SH. Viscoelastic properties of alkoxy silaneepoxy interpenetrating networks. Int J Adhes Adhes 2006;26:10315.
[109] Laly AP, Sabu T. Polarity parameters and dynamic mechanical behaviours of
chemically modied banana ber reinforced polyester composites. Compos
Sci Technol 2003;63:123140.
[110] Xie Y. Silane treated wood our as reinforcement for thermoplastic
composites. Project progressive report of German academic exchange
service, Gerog-August-University Goettingen, Germany; 2007.
[111] Tze WTY, Gardner DJ, Tripp CP, ONeill SC. Cellulosic ber/polymer adhesion:
effect of ber/matrix interfacial chemistry on the micromechanics of the
interphase. J Adhes Sci Technol 2006;20:164968.
[112] Le Digabel F, Boquillon N, Dole P, Monties B, Averous L. Properties of
thermoplastic composites based on wheatstraw lignocellulosic llers. J Appl
Polym Sci 2004;93:42836.
[113] Beshay A, Hoa SV. Improved interface bonding between cellulosic bers and
thermoplastics. Sci Eng Compos Mater 1992;2(2):8597.
[114] Hughes M, Carpenter J, Hill CAS. Deformation and fracture behaviour of ax
bre reinforced thermosetting polymer matrix composites. J Mater Sci
2007;42:2499511.
[115] Xanthos M. Processing conditions and coupling agent effects in
polypropylene/wood our composites. Plast Rubber Proc Appl 1983;3:2238.
[116] Gaiolas C, Costa AP, Nunes M, Santos Siva MJ, Belgacem MN. Grafting of paper
by silane coupling agents using cold-plasma discharge. Plasma Process Polym
2008;5:44452.
[117] Kim H, Jang J. Synthesis and characterization of vinyl silane modied
imidazole copolymer as a novel corrosion inhibitor. Polym Bull 1997;38:
24956.
[118] Shah GB, Fuzail M, Anwar J. Aspects of the crosslinking of polyethylene with
vinyl silane. J Appl Polym Sci 2004;92:3796803.
[119] Shieh YT, Tsai TH. Silane grafting reactions of low-density polyethylene. J
Appl Polym Sci 1998;69:25561.
[120] Arbelaiz A, Fernndez B, Cantero G, Llano-Ponte R, Valea A, Mondragon I.
Mechanical properties of ax bre/polypropylene composites. Inuence of
bre/matrix modication and glass bre hybridization. Compos Part A Appl
Sci 2005;36:163744.
[121] Bengtsson M, Oksman K. Prole extrusion and mechanical properties of
crosslinked woodthermoplastic composites. Polym Compos 2006;27: 18494.
[122] Bengtsson M, Oksman K. The use of silane technology in crosslinking
polyethylene/wood our composites. Compos Part A Appl Sci
2006;37:75265.
[123] Wong WK, Varrall DC. Role of molecular structure on the silane crosslinking
of polyethylene: the importance of resin molecular structure change during
silane grafting. Polymer 1994;35:544752.
[124] Herrera-Franco PJ, Valadez-Gonzlez A. Mechanical properties of continuous
natural bre-reinforced polymer composites. Compos Part A Appl Sci
2004;35:33945.
[125] Marsden JG, Orenski PJ. Azido-silane compositions. United States patent
3944574; 1976.
[126] Arenas JF, Marcos JI, Otero JC, Tocn IL, Soto J. Nitrenes as intermediates in
the thermal decomposition of aliphatic azides. Int J Quantum Chem 2001;84:
2418.
[127] Towo AN, Ansell MP. Fatigue evaluation and dynamic mechanical thermal
analysis of sisal brethermosetting resin composites. Compos Sci Technol
2008;68:92532.
[128] Kim HJ, Seo DW. Effect of water absorption fatigue on mechanical properties
of sisal textile-reinforced composites. Int J Fatigue 2006;28:130714.
[129] Baley C, Busnel F, Grohens Y, Sire O. Inuence of chemical treatments on
surface properties and adhesion of ax brepolyester resin. Compos Part A
Appl Sci 2006;37:162637.
[130] Hughes M, Hill CAS, Hague JRB. The fracture toughness of bast bre
reinforced polyester composites Part 1: evaluation and analysis. J Mater Sci
2002;37:466976.
[131] Serier A, Pascault JP, My LT. Reaction in aminosilaneepoxy prepolymer
systems I. Kinetics of epoxyamine reactions. J Polym Sci Polym Chem
1991;29:20918.
[132] Chiang CH, Koenig JL. Chemical reactions occurring at the interface of epoxy
matrix and aminosilane coupling agents in ber-reinforced composites.
Polym Compos 1980;1:8892.
[133] George J, Ivens J, Verpoest I. Mechanical properties of ax bre reinforced
epoxy composites. Die Angew Makromol Chem 1999;272:415.
[134] Uma Devi L, Bhagawan SS, Thomas S. Mechanical properties of pineapple leaf
ber-reinforced polyester composites. J Appl Polym Sci 1997;64: 173948.
[135] Ganan P, Mondragon I. Fique ber-reinforced polyester composites: effects of
ber surface treatments on mechanical behavior. J Mater Sci 2004;39:
31218.
[136] Pothan LA, Thomas S. Polarity parameters and dynamic mechanical
behaviour of chemically modied banana ber reinforced polyester
composites. Compos Sci Technol 2003;63:123140.
[137] ASTM D1037-99. Standard test methods for evaluating properties of woodbase ber and particle panel materials.

Y. Xie et al. / Composites: Part A 41 (2010) 806819


[138] Wolcott M. The role of thermoplastics in conventional wood composites. In:
Proceedings
30th
international
particleboard/composite
materials
symposium. Washington State University;1996.
[139] Lundin T, Falk RH, Felton C. Accelerated weathering of natural ber
thermoplastic composites: effects of ultraviolet exposure on bending
strength and stiffness. In: The sixth international conference on woodber
plastic composites, Madison, WI; 2001.
[140] Wallenberger FT, Weston N. Natural bers, plastics and composites. Kluwer
Academic Publishers; 2004.
[141] Mohanty AK, Misra M, Drzal LT. Natural bers, biopolymers, and
biocomposites. Taylor & Francis; 2005.
[142] Schirp A, Wolcott MP. Fungal degradation of woodplastic composites and
evaluation using dynamic mechanical analysis. J Appl Polym Sci
2006;99:313846.
[143] Morris PI, Cooper PA. Recycled plastic/wood composite lumber attacked by
Fungi. Forest Prod J 1997;48:868.

819

[144] Lomel-Ramrez MG, Ochoa-Ruiz HG, Fuentes-Talavera FJ, Garca-Enriquez S,


Cerpa-Gallegos MA, Silva-Guzmn JA. Evaluation of accelerated decay of
wood plastic composites by Xylophagus fungi. Int Biodeter Biodegr
2009;63:10305.
[145] Gnatowski M. Water absorption by wood plastic composites: eld and
laboratory challenges. In: The 10th international conference on progress in
biobre plastic composites, Toronto, Ontario; 2008.
[146] Sreekala MS, Thomas S. Effect of bre surface modication on water-sorption
characteristics of oil palm bres. Compos Sci Technol 2003;63:8619.
[147] Ibach RE, Clemons CM. Biological resistance of polyethylene composites
made with chemically modied ber or our. In: Proceedings of the 6th
pacic rim bio-based composites symposium, Portland, OR, USA, 2002. p.
57484.
[148] Hill CAS, Abdul Khalil HPS. The effect of environmental exposure upon the
mechanical properties of coir and oil palm ber reinforced composites. J Appl
Polym Sci 2000;77:132230.

You might also like