You are on page 1of 8

This article appeared in a journal published by Elsevier.

The attached
copy is furnished to the author for internal non-commercial research
and education use, including for instruction at the authors institution
and sharing with colleagues.
Other uses, including reproduction and distribution, or selling or
licensing copies, or posting to personal, institutional or third party
websites are prohibited.
In most cases authors are permitted to post their version of the
article (e.g. in Word or Tex form) to their personal website or
institutional repository. Authors requiring further information
regarding Elsevier’s archiving and manuscript policies are
encouraged to visit:
http://www.elsevier.com/copyright
Author's personal copy

Composites: Part A 40 (2009) 1975–1981

Contents lists available at ScienceDirect

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

Effects of fiber characteristics on the physical and mechanical properties


of wood plastic composites
Hassine Bouafif a, Ahmed Koubaa a,*, Patrick Perré b, Alain Cloutier c
a
Canada Research Chair on Wood Development, Characterization and Processing, Université du Québec en Abitibi-Témiscamingue, 445 BD de l’Université, Rouyn-Noranda, QC,
Canada J9X5E4
b
AgroParisTech, INRA, UMR1093, LERMAB, 14 rue Girardet, 54000 Nancy, France
c
Centre de Recherche sur le Bois, Université Laval, Pavillon Eugène-Kruger, QC, Canada G1K7P4

a r t i c l e i n f o a b s t r a c t

Article history: We investigated the effects of fiber variability, size, and content on selected mechanical and physical
Received 15 December 2008 properties of wood plastic composites. HDPE and fibers were compounded into pellets by twin-screw
Received in revised form 3 June 2009 extrusion and test specimens were prepared by injection molding. All tested properties vary significantly
Accepted 4 June 2009
with fiber origin. Higher fiber size produces higher strength and elasticity but lower energy to break and
elongation. The effect of fiber size on water uptake is minimal. Increasing fiber load improves the strength
and stiffness of the composite but decreases elongation and energy to break. Water uptake increases with
Keywords:
increasing fiber content.
A. PMCs
A. Wood
Ó 2009 Elsevier Ltd. All rights reserved.
B. Mechanical properties
B. Chemical properties

1. Introduction wood fibers to composite materials. They observed a correlation


between lignin content and longitudinal Young’s modulus, and
The effective use of wood-based particles and fibers as fillers or an optimal lignin content range for maximum fiber stiffness was
reinforcements in thermoplastic composites requires a fundamen- recorded for softwood Kraft fibers. Several attempts have been
tal understanding of the structural and chemical characteristics of made to correlate wood-based particles and fiber properties to
wood [1]. English and Falk [2] provide a comprehensive overview WPC properties [9–11]. A high aspect ratio (length/width) is very
of the factors that affect the properties of wood–plastic composites important in fiber reinforced composites, as it indicates potential
(WPC). Although several studies have shown that fiber-polymer strength properties [12]. Stark and Berger [11] investigated the ef-
compatibility can be enhanced by selecting suitable coupling fects of particle size on the properties of polypropylene filled with
agents [3,4], compatibility between polar wood fiber and non-polar wood flour. They concluded that melt flow index, heat deflection
thermoplastics remains key to extending the application limits of temperature, notched impact energy, and flexural and tensile mod-
the resultant composites [5]. Another, frequently cited key factor ulus and strength increase with increasing particle size. Later, Stark
in natural fiber thermoplastic composites is thermal degradation and Rowlands [10] reported that aspect ratio, rather than particle
[6]. Furthermore, different wood species have different anatomical size, has the greatest effect on strength and stiffness. They
structures. These structural differences govern the use of these suggested that particle size does not affect specific gravity.
materials in WPC. For example, fiber dimensions, strength, vari- Fiber content is an influential factor in WPC processing and
ability, and structure are important considerations. Maldas et al. properties. Zhang et al. [13] investigated the effects of fiber content
[7] are among the few researchers who have investigated the effect on mixing torque and rheological properties. They concluded that
of wood species on the mechanical properties of wood/thermo- increased wood fiber content results in increased steady state
plastic composites. They reported that differences in morphology, torque and viscosity. Lu and coworkers [14] concluded that the
density, and aspect ratios across wood species account for varying mechanical properties of the resultant WPC increase only at low
reinforcement properties in thermoplastic composites. Recently, weight percentages of wood filler. They found that tensile and
Neagu et al. [8] investigated the stiffness contribution of various flexural strengths reach a maximum at 15 wt% and 35 wt% wood
particle contents, respectively, and gradually decrease with a
further increase in wood particle content. Dänyädi et al. [15]
* Corresponding author. Tel.: +1 8197620971.
E-mail addresses: ahmed.koubaa@uqat.ca (A. Koubaa), perre@nancy-engref. inra.fr
reported that, at large wood content, considerable particle
(P. Perré), alain.cloutier@sbf.ulaval.ca (A. Cloutier). aggregation takes place, leading to lower strength due to the filler’s

1359-835X/$ - see front matter Ó 2009 Elsevier Ltd. All rights reserved.
doi:10.1016/j.compositesa.2009.06.003
Author's personal copy

1976 H. Bouafif et al. / Composites: Part A 40 (2009) 1975–1981

Table 1 Table 2
Hammer-milled particle classes. Injection molding machine settings.

Mean class (mesh) Size class interval Mold temperature – fixed/mobile 38 °C/38 °C
Injection pressure 900 kPa
Mesh lm
Injection pressure time 10 s
24 [20,28] [850, 600] Hold pressure 900 kPa
42 [35,48] [425, 300] Hold pressure time 4s
65 [48,100] [300, 150] Barrel temperature profile – feed, zone1, zone2, 160 °C, 190 °C, 190 °C,
nozzle 190 °C
Screw speed 135 rpm
Cooling time 15 s

failure to sustain the stress transferred from the polymer to the


matrix.
We investigated the effects of wood species, particle size, and tion molding machine in the conditions shown in Table 2. Speci-
fiber content on selected physical and mechanical properties of mens were made according to ASTM specifications for tensile,
wood particle-reinforced high-density polyethylene (HDPE). impact, and bending strength testing.
Specimens were stored in controlled conditions (50% relative
2. Materials and methods humidity and 23 °C) for 40 h prior to testing. Tensile and bending
tests were performed according to ASTM D638 [16] and ASTM
Five types of wood sawdust were investigated in this study: D790 [17], respectively. Energy to break is defined as the energy
eastern white cedar, with sapwood and heartwood sawdust trea- equivalent to the area under the tensile stress–strain curve.
ted separately; jack pine sawdust divided into wood sawdust and Composites were immersed in water at room temperature for
bark shavings; and black spruce sawdust. Sawdust and shavings two months. Specimens were weighed at regular intervals using
were hammer-milled into particles. Particles were then screened an analytical balance and water uptake was calculated.
and classified into three mesh sizes (Table 1) using an oscillating Experiments were conducted according to an incomplete facto-
multideck screen classifier. At this stage, the particles had a mois- rial design. As shown in Table 3, the factors are Species (five differ-
ture content of 10.5%. ent fibers), Size (three size classes), and Load (three contents).
The HDPE polymer (Goodfellow Corp., USA) used was a semi- Statistical analyses were conducted using a linear univariate proce-
crystalline polymer (70–80%) with 0.95 density, 9.0 g/10 min melt dure in SPSS (5% significance).
index, and 135 °C melting point. Ethylene maleic anhydride
copolymer (MAPE, A-CÒ 575A, Honeywell Int., USA) was used as 3. Results and discussion
a coupling agent.
WPC were produced in a two-stage process. In the first stage, Results of the analysis of variance (ANOVA) are shown in Table
wood particles were compounded into pellets at 25%, 35%, and 4. The ANOVA shows that all mechanical properties vary signifi-
45% by weight with the HDPE using a co-rotating twin-screw ex- cantly with fiber type, size, and content. Water uptake varies sig-
truder. Barrel temperatures of the four zones ranged from 180 °C nificantly with fiber type and content. The effect of fiber size on
to 190 °C from feeding to die zones. Screw speed was 240 rpm water uptake is not significant, but the interaction between fiber
and melt pressure at the die varied from 15 bar to 25 bar, depend- length and fiber content is highly significant. This result suggests
ing on wood particle content. In the second stage, injection WPC that the effect of fiber size on water uptake is dependent on fiber
test specimens were produced using a reciprocating screw injec- content.

Table 3
Factor combinations for the incomplete factorial design.

Species Cedar sapwood Cedar sapwood Jack pine bark Jack pine Black spruce
Size (mesh) 24 42 65 24 42 65 24 42 65 24 42 65 24 42 65
Load (%) 25 25 25 25 25 25 25 25 25 25 – – – 25 –
35 35 – 35 35 35 35 35 35 – 35 – 35 – –
45 45 – 45 45 45 45 45 45 – – 45 – – 45

Table 4
Results of the analysis of variance (F values) for selected mechanical and physical properties of wood plastic composites.

Source Mechanical properties Physical properties


Tensile Flexure Water uptake
r MOE e Energy to break MOR MOE e 48 h 4 weeks
Model 96.1b 34.0b 310.7b 186.0b 254.3b 294.2b 201.6b 3.3b 13.3b
Fiber 300.7b 76.4b 279.1b 476.6b 631.5b 667.5b 108.9b 1.7 ns 7.9b
Size 224.6b 27.8b 32.1b 29.2b 496.4b 410.3b 117.0b 2.1 ns 0.6ns
Content 56.7b 166.4b 2214.1b 747.3b 548.7b 1542.1b 1458.8b 17.8b 143.5b
Fiber  size 72.6b 7.5b 252.4b 183.4b 115.7b 108.5b 116.1b 1.7 ns 1.3ns
Fiber  content 12.3b 19.2b 65.4b 22.0b 74.0b 98.7b 44.7b 4.0b 1.8ns
Size  content 16.7b 8.1b 20.3b 16.3b 88.2b 69.5a 4.8b 1.4 ns 3.3a
Fiber  size  content 34.0b 6.1b 14.8b 21.1b 38.5a 32.6b 7.4b 3.5b 3.6b
a
Significant at 0.05; ns: non-significant at 0.05.
b
Significant at 0.01.
Author's personal copy

H. Bouafif et al. / Composites: Part A 40 (2009) 1975–1981 1977

2.8 equal (0.28), whereas jack pine fibers show the most oxidized sur-
a Flexural Modulus of Elasticity
Tensile Modulus of Elasticity
face (0.35). Bark fibers show the lowest O/C ratio (0.18). Fig. 2a
2.4
shows close relationships between atomic O/C ratio and tensile
Modulus of Elasticity (GPa)

and flexural strength. Similar tendencies are obtained for the ten-
sile and flexural modulus of elasticity (not shown). The assumed
2.0 association of low O/C ratio with a lignin-rich surface [18], which
prevents the formation of ester bonds between fibers and the cou-
1.6 pling agent (MAPE), could explain the lower strength values ob-
tained for the bark composites.
For composites made with semi-crystalline matrix polymers,
1.2
crystallinity is another important factor in determining the stiff-
ness and fracture behavior of the crystallized matrix polymer
0.8 [19]. Crystallinity depends on the processing parameters, e.g., crys-
tallinity temperature, cooling rate, nucleation density, annealing
HDPE Neat Cedar (Heartwood) Jack pine time, and fiber type [20,21]. Bouafif et al. [22] investigated the effi-
Cedar (Sapwood) Bark Black spruce ciency of the same wood filler as a nucleating agent using non-iso-
thermal analysis. Nucleating activity was determined for various
50
Maximum Flexural Strength wood fillers by the crystalline weight fraction of the HDPE matrix.
b Maximum Tensile Strength
45

40
Strength (MPa)

35

30

25

20

15
HDPE Neat Cedar (Heartwood) Jack pine
Cedar (Sapwood) Bark Black spruce

Fig. 1. Variation in (a) tensile and flexural modulus of elasticity and maximum
flexural and tensile strength for wood particle/HDPE injection-molded composites
with different filler species at constant fiber load (35%) and particle size (42 mesh).

3.1. Effect of fiber type

The mechanical proprieties of the injected wood plastic com-


posites vary significantly with fiber type (Table 4). Fig. 1a and b
illustrates the tensile and flexural modulus of elasticity and maxi-
mum strength, respectively, of wood particle-reinforced HDPE
composites made with various wood types. WPC made with jack
pine and black spruce particles exhibit the highest modulus of elas-
ticity and strength, whereas bark and eastern white cedar WPC
show the lowest properties. Average flexural modulus of elasticity
ranges from 2.3 GPa to 2.2 GPa for jack pine and black spruce WPC,
respectively, while maximum flexural strength is approximately
equal (40 MPa) for both species. In other words, both modulus of
elasticity and maximum strength of HDPE are enhanced at least
1.5 times when jack pine or black spruce is added.
In a previous investigation, Bouafif et al. [18] used the atomic
concentration ratio of oxygen to carbon (O/C) on the fiber surface
as an initial indication of surface oxidation (Table 5). The O/C Fig. 2. Variation in tensile and flexural strength with (a) O/C ratio and (b) relative
atomic ratios for eastern white cedar and black spruce fibers are crystallinity.

Table 5
Physical and chemical properties of wood fibers.

Species Specific gravity Extractives (%) Lignin (%) Carbohydrates (%) O/Ca atomic ratio Composite crystallinity (%)b
White cedar 0.32 12 25 67 0.28 76.0
Jack pine 0.43 3 27 70 0.35 77.4
Black spruce 0.46 3 25 72 0.28 76.2
Jack pine bark 0.40 10 48 42 0.18 74.6
a
Data from Bouafif et al. [18].
b
Data from Bouafif et al. [22]
Author's personal copy

1978 H. Bouafif et al. / Composites: Part A 40 (2009) 1975–1981

When tensile and flexural strength were plotted against the crys- [18]. To achieve optimal strength of the WPC, black spruce or jack
talline weight fraction (Fig. 2b), an affinity relationship was estab- pine particles are the most appropriate for reinforcements.
lished. Other factors could also explain this difference, including Water uptake of the injected WPC varies among the five studied
the lower intrinsic fiber strength of bark fibers compared to wood fibers. Bark WPC exhibit lower water absorption compared to those
fibers. made with wood particles. This result could be explained by the
Fig. 3a and b depicts the energy to break and elongation at highly different chemical composition of bark fibers compared to
break, respectively, of the obtained composite materials. Compos- wood fibers (Table 4). Because bark contains higher hydrophobic
ites filled with eastern white cedar sapwood and bark show brittle volume content (lignin and extractives) and lower hydrophilic con-
behavior, with 90% and 60% lower energy to break than black tent (cellulose and hemicelluloses), it would be expected to show
spruce. This result is mainly explained by poor adhesion between lower water uptake compared to wood fibers (Fig. 4). Nevertheless,
the particles of some species and the HDPE matrix. Because white high among-species variation is found in the wood–fiber WPC for
cedar wood and bark contain high amounts of extractives at the water uptake. These differences could be explained by several fac-
surface compared to jack pine and black spruce, a weak surface tors, including wood chemical composition. For example, eastern
boundary layer can be formed, making the coupling agent less white cedar heartwood has higher extractive content than its sap-
effective in forming a cross-linking network with the cellulose wood. Thus, as expected, WPC made with heartwood shows lower
[23]. This result is observed for bark-particle WPC, which shows water uptake than WPC made with sapwood (Fig. 4).
the second highest elongation at break after black spruce WPC.
This could be explained by the good dispersion of this particle type. 3.2. Effect of particle size
Thus, some lipophilic extractives might help to disperse the parti-
cles during WPC preparation. These results agree with Maldas et al. The effect of particle size on the mechanical properties investi-
[7], who reported that differences in morphology, density, and as- gated in this study is highly significant (Table 4). In general,
pect ratios across wood species account for varying reinforcement increasing fiber size improves the modulus of elasticity and maxi-
in thermoplastic composites. It is therefore important to select the mum strength in both tensile and flexure tests (Fig. 5a–d). This re-
appropriate wood species for optimal WPC end use. Accordingly, sult is consistent with previous reports on wood-particle
depending on the intrinsic characteristics of wood fibers, the prop- thermoplastic composites [10,11,24,25]. As for tensile modulus,
erties of the resultant WPC differ to a large extent. The order of the flexural modulus of elasticity shows a steady increase with
reinforcement performance of the investigated fibers closely fol- increasing particle size at higher filler content (Fig. 5c). It rises from
lows the rank of their ability to form ester bonds with MAPE 2.1 GPa at particle sizes ranging from 100 mesh to 48 mesh (150–
300 lm) to 2.7 GPa at average particle size of 24 mesh (710 lm)
and 45 wt% filler content. These results are in good agreement with
previously reported data [11,24].
220
Flexural strength development (Fig. 5d) also demonstrates that
a particle size has greater influence at higher fiber load (45 wt%),
200 with approximately 24% higher strength when average particle
Energy to Break (J/m 2 X103 )

size increases from 65 mesh to 24 mesh. On the other hand, the


180 incorporation of wood particles in the HDPE matrix steadily in-
creases tensile strength (Fig. 1b), independently of filler content.
160
When average particle size increases from 65 mesh (230 lm) to
140
24 mesh (710 lm), tensile strength improves by 43%, 10%, and
12% at 25 wt%, 35 wt%, and 45 wt% filler content, respectively. Zaini
120 et al. [24] reported increasing maximum tensile strength with
increasing particle size for isotactic polypropylene filled with 250
100 mesh to 63 mesh oil palm wood flour, while Stark and Rowlands
[10] recommended the use of higher aspect ratio wood fibers,
80
Cedar (Sapwood) Bark Black spruce rather than larger size fibers, to increase WPC strength. However,
Cedar (Heartwood) Jack pine Migneault et al. [26] showed that maintaining constant fiber diam-
eter and increasing fiber length improves the strength and elastic-
8.0 ity of HDPE-filled composites. Stark and Berger [11] found that
b larger filler particles (greater than 250 lm) decrease tensile
7.5
properties.
Elongation at Break (%)

7.0

6.5

6.0

5.5

5.0

4.5

4.0
Cedar(Sapwood) Bark Black spruce
Cedar(Heartwood) Jack pine

Fig. 3. Variation in (a) energy to break and (b) maximum tensile elongation at break
for wood particle/HDPE injection-molded composites with different filler species at Fig. 4. Average water uptake in composite materials with various wood fibers after
constant load (35%) and particle size (42 mesh). four weeks of water immersion.
Author's personal copy

H. Bouafif et al. / Composites: Part A 40 (2009) 1975–1981 1979

32
25 wt% 25 wt%
a 35 wt% b 35 wt%
Tensile Modulus of Elasticity (GPa)

Maximum Tensile Strength (MPa)


45 wt% 45 wt%
2.0 28

24
1.6

20

1.2
16
65 42 24 65 42 24
Average Particle Size (mesh) Average Particle Size (mesh)

3.2 50
25 wt% 25 wt%
c 35 wt% d
Flexural Modulus of Elasticity (GPa)

35 wt%

Maximum Flexural Strength (MPa)


2.8 45 wt% 45 45 wt%

2.4 40

35
2.0

30
1.6

25
1.2

20
65 42 24 65 42 24
Average Particle Size (mesh) Average Particle Size (mesh)

Fig. 5. Effect of wood particle size with various filler contents on (a) tensile modulus of elasticity (b), tensile strength, (c) flexural modulus of elasticity, and (d) flexural
strength of WPC.

Fig. 6a and b shows the energy to break and elongation at break around the wood particles. These voids in the bulk matrix are read-
of the resultant composite materials, respectively. Both properties ily filled with water. At higher contents, the effect of fiber size is
decrease with larger particle size. The effect is more pronounced as not significant. This could be attributed to the fact that, at higher
particle concentration increases. Despite the significant increase in contents, the effect of fiber size is minimal compared to that of
tensile strength with increasing particle size, the rupture energy content. In addition, fiber size changes during processing. We
decreases (Fig 6a), which could be explained by the decrease in investigated this development, mainly after compounding (results
elongation at break. This is a common tendency that has been re- not shown), and observed that particles with higher aspect ratios
ported with inorganic filler as well [27]. Composite strength de- are subject to severe damage due to higher shear stresses that
pends on the debonding process at the fiber end and the fiber’s develops in the conical counter-rotating intermeshing twin-screw
pull-out process during interface failure [28]. The phenomenon is extruder during mixing and compounding. Apparently, particle
more marked as particle size increases. Because cracks travel length distribution after compounding skews towards the shortest
around the wood particles, the fracture surface area increases with particle length.
increasing particle size. As a result, less energy is required to frac- WPC water absorption tends to increase up to 8 weeks of
ture a specimen containing larger particles [11]. immersion, as shown in Fig. 8. This negates the argument that
In summary, although incorporating larger particle sizes into WPC reaches saturation at short and medium immersion times.
the HDPE matrix effectively improves strength properties, this For example, assuming that moisture is absorbed by wood particles
improvement comes at the price of energy to break. Thus, depend- alone, then the average moisture content of wood particles is 10.4%
ing on end use, the composite should be optimized for either stiff- in composites filled with 42 mesh particles at 35 wt%. However,
ness or energy to break, by adjusting both filler particle size and when neat wood particles are immersed in water, moisture con-
concentration. tent readily exceeds 30%, or fiber saturation. Therefore, WPC water
The effect of particle size on water uptake is not significant, as uptake may actually be slowed down rather than delayed.
shown in Table 4. However, the interaction between fiber size
and fiber content is significant. Thus, the effect of fiber length on 3.3. Effect of fiber content
water uptake is dependent on fiber content. At lower contents,
the larger the particle size, the higher the water absorption (Fig As shown in Table 4, the effect of fiber content is highly signif-
7), which is in good agreement with previous reports [29,30]. This icant for all tested properties. In general, high wood particle con-
can be explained in two ways: (i) larger particles lead to greater tent yields materials with high rigidity and strength. Fig. 8a and
hydrophilic exposed surfaces; and (ii) poor adhesion between b illustrates the effect of fiber content on rigidity and strength
wood particles and the polymer matrix generates void spaces properties, respectively, in eastern white cedar particle-based
Author's personal copy

1980 H. Bouafif et al. / Composites: Part A 40 (2009) 1975–1981

280 3.2
25 wt% MOEF
a 35 wt% MOET
45 wt% 2.8
240

Modulus of elasticity (GPa)


Energy to Break (J/m2 x103 )

2.4
200

2.0
160
1.6

120
1.2

80 0.8

0 wt % (Neat HDPE) 25 wt % 35 wt % 45 wt %
40
65 42 24 Wood particle content
Average Particle Size (mesh)
60
Fmax.
16
Rm
25 wt%
b 35 wt%
50

45 wt%
Elongation at Break (%)

Strength (MPa)
12 40

30

8
20

10
4
0 wt % (Neat HDPE) 25 wt % 35 wt % 45 wt %

65 42 24 Wood particle content

Average Particle Size (mesh) Fig. 8. Variation in tensile and flexural (a) modulus of elasticity, and (b) maximum
strength of eastern white cedar–HDPE composites.
Fig. 6. Effect of wood particle size and filler content on (a) energy to break (tensile
energy) and (b) tensile elongation at break.

creases from 25 wt% to 45 wt%. However, wood fibers cause a dra-


composites. Tensile modulus of elasticity (TMOE) of the resultant matic decrease in elongation at break after flexion and tensile
composites increases steadily with wood fiber content, from stress, as shown in Fig. 9a. Over 400% reduction in elongation at
0.85 GPa for neat HDPE to 1.75 GPa for WPC with 45 wt% wood break (e RB) is recorded at 45% particle content compared to neat
content. Loading neat HDPE with 45 wt% wood fibers increases HDPE. These results are in good agreement with previous reports
the flexural modulus of elasticity by over three times. on wood-particle thermoplastic composites [31,32]. Over 135%
Compared to neat HDPE, maximum flexural (Fmax) and maxi- reduction in energy to break is recorded at 45% particle content
mum tensile strength (Rm) of HDPE nearly double with the addi- compared to 25% wood-based composites (Fig. 9b).This reduction
tion of 45% wood particles. Flexural and tensile strength at break, in elongation is expected, given that adding fiber changes the vis-
respectively, improve by 53% and 18% when wood filler content in- coelastic behavior of HDPE to a more ductile behavior.

2.5 4. Conclusions
20 - 28 mesh
35 - 48 mesh
48 - 100 mesh
Wood fiber variability has a significant effect on the mechanical
2.0 and physical properties studied. This effect is explained by varia-
tions in fiber surface properties, relative crystallinity, thermal deg-
Water Uptake (wt %)

radation, and wood specific gravity. Composites made with bark


1.5
particles exhibit lower water absorption compared to those made
with wood particles. Differences in chemical composition between
1.0 bark and wood are among the plausible explanations for the differ-
ences in water uptake. Increasing fiber size improves strength and
stiffness, but reduces elongation and energy to break. The effect of
0.5 fiber size on water uptake is minimal compared to the effects of fi-
ber origin and content. Increasing fiber content improves the
strength and stiffness of composites, but decreases elongation
0.0
0 200 400 600 800 1000 1200 1400 and energy to break. Water uptake increases substantially with
Immersion Time (h) increasing fiber content.
Composites made with bark particles exhibit lower water
Fig. 7. Average water uptake in composite materials with wood particle size. absorption compared to those made with wood particles. Differ-
Author's personal copy

H. Bouafif et al. / Composites: Part A 40 (2009) 1975–1981 1981

12 [4] Maldas D, Kokta BV, Raj RG, Daneault C. Improvement of the mechanical
Rm properties of sawdust wood fibre–polystyrene composites by chemical
a RB treatment. Polymer 1988;29(7):1255–65.
10 [5] Zadorecki P, Michell AJ. Future prospects for wood cellulose as reinforcement
Elongation at max. strength (%)

in organic polymer composites. Polym Compos 1989;10(2):69–77.


[6] Burgstaller C. Processing of thermal sensitive materials—a case study for wood
Elongation at break (%)

8 plastic composites. Monatsh Chem 2007;138(4):343–6.


[7] Maldas D, Kokta BV, Daneault C. Thermoplastic composites of polystyrene:
effect of different wood species on mechanical properties. J Appl Polym Sci
6
1989;38(3):413–39.
[8] Neagu RC, Gamstedt EK, Berthold F. Stiffness contribution of various wood
fibers to composite materials. J Compos Mater 2006;40(8):663–99.
[9] Bledzki AK, Faruk O. Microcellular injection molded wood fiber–PP
4
composites: part II–Effect of wood fiber length and content on cell
morphology and physico-mechanical properties. J Cell Plast 2006;42(1):77–88.
[10] Stark NM, Rowlands RE. Effects of wood fiber characteristics on mechanical
2
properties of wood/polypropylene composites. Wood Fiber Sci 2003;35(2):167–74.
[11] Stark NM, Berger MJ. Effect of particle size on properties of wood–flour
reinforced polypropylene composites. In: Fourth international conference on
0 woodfiber–plastic composites. (Madison WI): Forest Product Society; 1997.
0 wt % (Neat HDPE) 25 wt % 35 wt % 45 wt % [12] Rowell RM, Han JS, Rowell JS. Characterization and factors effecting fiber
wood particle content properties. In: Natural polymers and agrofibers composites, Sãn Carlos
(Brazil); 2000. p. 292.
200 [13] Zhang SY, Zhang Y, Bousmina M, Sain M, Choi P. Effects of raw fiber materials,
fiber content, and coupling agent content on selected properties of
b polyethylene/wood fiber composites. Polym Eng Sci 2007;47(10):1678–87.
180 [14] Lu JZ, Wu Q, Negulescu II. Wood-fiber/high-density-polyethylene composites:
Energy to Break (x 10 3 J/m²)

coupling agent performance. J Appl Polym Sci 2005;96(1):93–102.


160 [15] Dänyädi L, Janecska T, Szabo Z, Nagy G, Möczo J, Pukänszky B. Wood flour filled
PP composites: compatibilization and adhesion. Compos Sci Technol
2007;67(13):2838–46.
140 [16] ASTM D638-03. Standard test method for tensile properties of plastics. West
Conshohocken: ASTM International; 2003.
[17] ASTM D790-03. Standard test method for flexural properties of unreinforced
120
and reinforced plastics and electrical insulating materials plastics. West
Conshohocken: ASTM International; 2003.
100 [18] Bouafif H, Koubaa A, Perré P, Cloutier A, Riedl B. Analysis of among-species
variability in wood fiber surface using DRIFTS and XPS: effects on esterification
efficiency. J Wood Chem Technol 2008;28(4):296–315.
80
[19] Joseph PV, Joseph K, Thomas S, Pillai CKS, Prasad VS, Groeninckx G, et al. The
thermal and crystallisation studies of short sisal fibre reinforced polypropylene
60 composites. Compos Part A Appl Sci Manufact 2003;34(3):253–66.
25 wt % 35 wt % 45 wt % [20] Wang C, Liu CR. Transcrystallization of polypropylene composites: nucleating
wood Particle content ability of fibres. Polymer 1999;40(2):289–98.
[21] Thomason JL, Van Rooyen AA. Transcrystallized interphase in thermoplastic
Fig. 9. Variation in tensile (a) elongation at maximum strength and (b) energy to composites – Part I Influence of fibre type and crystallization temperature. J
break of eastern white cedar–HDPE composites. Mater Sci 1992;27(4):889–96.
[22] Bouafif H, Koubaa A, Perré P, Cloutier A, Riedl B. Wood particle/HDPE
composites: thermal sensitivity and nucleating ability of wood particles. J
Appl Polym Sci 2009;113 (1): 593–600.
ences in chemical composition between bark and wood are among [23] Saputra H, Simonsen J, Li K. Effect of extractives on the flexural properties of
the plausible explanations for the differences in water uptake. wood/plastic composites. Compos Interface 2004;11(7):515–24.
[24] Zaini MJ, Fuad MYA, Ismail Z, Mansor MS, Mustafah J. The effect of filler
content and size on the mechanical properties of polypropylene/oil palm wood
Acknowledgments flour composites. Polym Int 1996;40(1):51–5.
[25] Dikobe DG, Luyt AS. Effect of filler content and size on the properties of
ethylene vinyl acetate copolymer–wood fiber composites. J Appl Polym Sci
The Authors are grateful to the Canada Research Chair Program; 2007;103(6):3645–54.
the Ministère du développement économique, de l’Innovation et de [26] Migneault S, Koubaa A, Erchiqui F, Chaala A, Englund K, Krause C, et al. Effect of
l’exportation du Québec (MDEIE); Caisse Populaire Desjardins; the fiber length on processing and properties of extruded wood-fiber/HDPE
composites. J Appl Polym Sci 2008;110(2):1085–92.
Centre de Technologie des Résidus Industriels (CTRI); Tembec; and [27] Dubnikova IL, Berezina SM, Antonov AV. Effect of rigid particle size on the
La Fondation de l’UQAT for financial support. toughness of filled polypropylene. J Appl Polym Sci 2004;94(5):1917–26.
[28] Liu YH, Xu JQ, Ding HJ. Mechanical behavior of a fiber end in short fiber
reinforced composites. Int J Eng Sci 1999;37(6):753–70.
References [29] Tajvidi M, Najafi SK, Moteei N. Long-term water uptake behavior of natural
fiber/polypropylene composites. J Appl Polym Sci 2006;99(5):2199–203.
[1] Stokke DD, Gardner DJ. Fundamental aspects of wood as a component of [30] Steckel V, Clemons CM, Thoemen H. Effects of material parameters on the
thermoplastic composites. J Vinyl Addit Technol 2003;9(2):96–104. diffusion and sorption properties of wood–flour/polypropylene composites. J
[2] English BW, Falk RH. Factors that affect the application of woodfiber–plastic Appl Polym Sci 2007;103(2):752–63.
composites. In: Proceedings of the woodfiber–plastic composites conference, [31] Robin JJ, Breton Y. Reinforcement of recycled polyethylene with wood fibers
Madison, Wisconsin; 1995. p. 183. heat treated. J Reinf Plast Compos 2001;20(14):1253–62.
[3] Bledzki AK, Reihmane S, Gassan J. Thermoplastics reinforced with wood fillers: [32] Park BD, Balatinecz JJ. Mechanical properties of wood-fiber/toughened
a literature review. Polym Plast Technol Eng 1998;37(4):451–68. isotactic polypropylene composites. Polym Compos 1997;18(1):79–89.

You might also like