Alkali Treatment of Hemp Fibres For The Production of Aligned Hemp Fibre Mats For Composite Reinforcement - Tom Sunny
Alkali Treatment of Hemp Fibres For The Production of Aligned Hemp Fibre Mats For Composite Reinforcement - Tom Sunny
8 Abstract
9 The main objective of this study was to produce aligned hemp fibre mats from high strength hemp
10 fibres using dynamic sheet forming (DSF). Alkali treatment of hemp fibre was carried out at
11 ambient and high temperature to separate fibres. Single fibre tensile testing was used to assess the
12 tensile properties of the fibres. It was found that the highest tensile properties were exhibited by
13 high temperature treated fibre, whereas the tensile properties exhibited by ambient temperature
14 treated fibre were lower than for untreated fibre. It was also found that fibre granulated after high
15 temperature treatment, was better separated than that granulated before high temperature
16 treatment. This well-separated fibre could successfully be formed into mats using DSF. The
17 orientation of the formed mat was analysed using ImageJ (NIH, USA) software by which the
18 potential of DSF to produce aligned hemp fibre mat was supported. The mechanical performance
19 of composite reinforced by these aligned hemp fibre mats were assessed.
20 Keywords: Dynamic sheet forming (DSF); Alkali treatment; Aligned fibre mats;
21 Hemp fibre
22 1. Introduction
23 A major area of recent technological development has been that of natural plant fibre composites
24 (NPFCs). Generally, NPFCs consist of plant fibres as reinforcement in a polymer matrix. Natural
25 plant fibres (NPFs) have undergone increased industrial uptake because of their favourable
26 characteristics such as lower density, higher specific strength and stiffness, lower cost, and with
27 lower associated hazards during manufacture (Pickering 2008, Kabir, Wang et al. 2012, Kabir,
28 Wang et al. 2013, Pickering, Efendy et al. 2015). Hemp fibres are attractive reinforcements for
29 NPFCs as they exhibit high tensile strength ranging between 550-1110 MPa (Beckermann and
30 Pickering 2008, Pickering, Efendy et al. 2015). Hemp fibres are also more environmentally
31 friendly than most fibres since they can be grown without pesticides and herbicides.
1
32 The main reasons that limit NPFCs in industry are moisture absorption, lower strength and greater
33 variability of properties compared to synthetic fibre composites (Pickering, Efendy et al. 2015).
34 Previous research has shown that chemical treatments of reinforcing fibres can help to overcome
35 these limitations (Mwaikambo and Ansell 2002, Li, Tabil et al. 2007, John and Anandjiwala 2008,
36 Kabir, Wang et al. 2012, Pickering, Efendy et al. 2015). Among the different chemical treatments,
37 alkali treatment with sodium hydroxide (NaOH) is the most widely used. This treatment removes
38 fibre constituents such as pectin, hemicellulose, lignin and waxes from NPFs bringing about fibre
39 separation and can enhance fibre properties (Beckermann and Pickering 2008, Islam, Pickering et
40 al. 2011, Pickering, Efendy et al. 2015). Modest treatments have been seen to bring about
42 aforementioned materials, whereas harsher treatments have been shown to convert crystalline
43 cellulose to amorphous cellulose and possibly resulted in chain scission (Sawpan, Pickering et al.
44 2011, Kabir, Wang et al. 2012). These alkali treatments have been carried out by different
45 researchers with varying process parameters including at ambient temperatures (AT) as well as at
46 high temperatures (HT) (Mwaikambo and Ansell 2002, Beckermann and Pickering 2008, Islam,
47 Pickering et al. 2010, Islam, Pickering et al. 2011, Sawpan, Pickering et al. 2011, Kabir, Wang et
48 al. 2012, Kabir, Wang et al. 2013, Efendy and Pickering 2014). Ambient temperature treatments
49 have many advantages such as simplicity, low cost and they can be easily carried out in large
50 volumes compared to HT treatment. However, there is limited information available on the effect
51 of alkali treatment temperature (high temperature versus ambient temperature) on the tensile
52 properties of individual hemp fibres. Also, where the tensile properties are given, these values are
53 based on variable testing regimes including: fibre bundles as opposed to single fibres and
54 unspecified gauge lengths (Beckermann 2007). Therefore, direct comparison is needed between
56 To encourage uptake of NPFCs, convenient forms of NPFs that could be used as alternatives to
57 synthetic fibre mats in standard moulding operations would be helpful. Dynamic sheet forming
58 (DSF) is a method that can be used to produce preferentially aligned short NPF mats (Sunny,
59 Pickering et al. 2017). Previous studies have demonstrated that these mats can be incorporated into
60 polymer matrices to produce composites, which exhibited higher tensile strength as well as
61 Young’s modulus compared to randomly oriented fibre mats composites (Pickering and Efendy
62 2016, Pickering and Le 2016, Sunny, Pickering et al. 2017). Prior to DSF, good fibre separation is
2
63 needed. Alkali treated hemp and harakeke fibres have been reported as used for the production of
64 DSF mats, (Le and Pickering 2015, Pickering, Efendy et al. 2015, Ghazali and Efendy 2016) but
65 the tensile strengths of these treated fibres were reduced compared to raw fibre due to the
66 weakening of structural components (Islam 2008, Ghazali and Efendy 2016, Le 2016, Pickering
67 and Efendy 2016). Therefore, further research is needed to develop processing parameters to
68 improve fibre separation without reduction in tensile properties of fibres prior to DSF. In the
69 current research, the alkali treatments used are modifications of selected alkali treatments from the
70 literature that report improved tensile strength for NPFs (Beckermann and Pickering 2008,
71 Oushabi, Sair et al. 2017). The data obtained for the tensile strengths of the hemp fibres were
73 2. Experimental
74 2.1 Materials
75 Industrial hemp fibre was obtained from Moffett Orchards Ltd., New Zealand. The bast fibres
76 were hand separated from the stalks. The chemicals used for the experiments were sodium
77 hydroxide (NaOH) and sodium sulphite (Na2SO3) supplied by Sigma Aldrich. Polypropylene (PP)
78 random copolymer SKRX3600 supplied by Clariant (New Zealand) Limited, with a melt index of
79 18 g/10min and a density of 0.9 g/cm3, was used as the matrix. The coupling agent used was A-C
80 950P maleic anhydride polypropylene (MAPP) supplied by Honeywell International Inc., USA
81 2.2 Methods
83 High temperature (HT) and ambient temperature (AT) alkali fibre treatments were carried out on
84 pre-dried hand separated hemp fibres. For ambient temperature treatment, fibres were granulated
85 using an 8 mm mesh in a laboratory scale Castin granulator and then immersed in 5wt% sodium
86 hydroxide (NaOH) solution in a glass beaker as shown in Fig. 1a, for one (AT/one hour) or two
87 hours (AT/ two hours) with a fibre to solution ratio of 1:8. The temperature inside and outside the
88 beaker was measured using a thermometer. The measured room temperature was between 20.5 °C
89 and 22 °C, whereas the temperature inside the beaker ranged from 30.5 °C at the start of the
3
90 treatment to 24 °C at the end of the treatment. For high temperature treatment, fibre and a solution
91 of 5wt% NaOH and 2wt% Na2SO3 with a fibre to solution ratio of 1:8 were placed in stainless
92 steel canisters (SSCs) Fig. 1b. These canisters were then positioned inside a laboratory scale pulp
94 which was set to operate with a time-temperature profile as shown in Fig. 2. Granulation was
95 either conducted before or after treatment. The fibres were washed after the treatments for about
a b
97
98 Fig. 1. (a) Set up used for the AT treatment (b) SSC and PID system used for the HT treatment.
99
100 Fig. 2. Time-temperature profile used for the HT treatment.
102 The ASTM D 3379-75: Standard Test Method for Tensile Strength and Young’s Modulus for
103 High-Modulus Single-Filament Materials (ASTM 1986) was followed to determine the tensile
104 strength and Young’s modulus of untreated (UT), AT/one hour treated, granulated before HT
105 treated and granulated after HT treated hemp fibres. The untreated fibres were soaked in water for
4
106 around 10 days to remove dirt from the fibre surface. Two-millimetre thick cardboard was used for
107 mounting tabs with a gauge length of two millimetres as schematically represented in Fig. 3.
108 Selected single fibres were adhered to the mounting tabs by the application of polyvinyl acetate
110
111 Fig. 3. Schematic representation of a mounting tab used for tensile testing of single fibres.
112 For the measurement of single fibre diameter, optical images were captured of single fibres as
113 shown in Fig. 4, by means of an Olympus BX60F5 optical microscope fitted with a Nikon camera.
114 The diameter was measured at five different points along each fibre (as hemp fibres have variable
115 diameters across their length) and average values were used for the calculations.
116
117 Fig. 4. Single hemp fibre as observed under optical microscope: (a) UT, (b) AT/one hour, (c)
118 granulated before HT and (d) granulated after HT treated hemp fibres.
5
119 The single fibres were then tensile tested using an Instron-4204 universal testing machine after
120 burning off the supporting sides using a hot wire cutter. The test was carried out at a rate of 0.5
121 mm/min with a 10 N-load cell. Thirty samples were tested for each treatment and system
122 compliance was determined experimentally to obtain Young’s moduli according to the procedure
123 described in ASTM D 3379-75. For this, mounting tabs with gauge lengths of 5, 10 and 15 mm
125 The Weibull distribution is commonly used to analyse the strength variation for natural fibres
126 (Biagiotti, Puglia et al. 2004, Pickering, Beckermann et al. 2007, Zafeiropoulos and Baillie 2007,
127 Efendy and Pickering 2014). Here, the rearranged two-parameter Weibull cumulative distribution
128 expression (Biagiotti, Puglia et al. 2004, Pickering, Beckermann et al. 2007), as shown below was
129 used to analyse data obtained for different single fibre testing statistically.
131 where w is the Weibull modulus (shape parameter) and σ0 characteristic strength (scale parameter).
132 These parameters are important as they describe the variability of the fibre failure strength
133 (Zafeiropoulos and Baillie 2007). A Weibull plot of lnln (1/(1-Pf) versus lnσ provides a straight
136 A Hitachi S-4100 SEM was used to obtain micrographs of fibres. Carbon tapes were employed to
137 mount the samples on aluminium stubs and were then sputter coated with platinum to make them
138 conductive.
140 To assess fibre crystallinity, a Philips X’Pert diffractometer fitted with a ceramic X-ray diffraction
141 tube was used. For the measurements, the fibres were chopped and pressed into a disk using a
142 cylindrical steel mould. The scanning range was between 5 o and 45o by employing CuKα radiation
143 (λ=1.54 nm) with a voltage and current of 45 mV and 40 mA respectively. Crystallinity index (Ic )
144 of the fibres was calculated using the Segal method (Segal, Creely et al. 1959):
6
146 where I200 is the maximum intensity of the (200) lattice diffraction peak at a 2θ angle of between
147 22o and 23o, and Iam is the minimum intensity of diffraction at an angle 2θ between 18 o and 19o
148 representing amorphous materials (Pickering, Beckermann et al. 2007, French 2014).
150 An FTIR Digilab FTS-40 was used to obtain infrared spectra of untreated, AT/one hour treated
151 and granulated after HT treated hemp fibres. Hemp fibre samples were ground to fine powder
152 using a Retsch MM400 ball mill. The ground powder for each sample was then mixed and
153 compressed with KBr (potassium bromide) using a hydraulic press by applying 8 tonnes/cm2
156 Thermal gravimetric analysis (TGA) of untreated, AT/one hour treated and granulated after HT
157 treated hemp fibres was carried out using a PerkinElmer simultaneous thermal analyser STA 800.
158 Data were obtained at a rate of 10 °C/min with a heating range of 40 °C to 500 °C and a static
161 The OrientationJ plugin of ImageJ was used to assess the orientation of fibres in fibre mats.
162 Availability and ease of use make ImageJ attractive (Abràmoff, Magalhães et al. 2004, Schneider,
163 Rasband et al. 2012). The principles behind this analysis tool are available in the literature
164 (Rezakhaniha, Agianniotis et al. 2012, Gesellchen, Bernassau et al. 2014, Shah, Vollrath et al.
165 2015, Püspöki, Storath et al. 2016, Sunny, Pickering et al. 2018). OrientationJ can provide a fibre
166 orientation distribution profile based on the orientation of fibres within a mat analysed. The peak
167 of the profile indicates the predominant orientation of fibres within the mat.
168 Additionally, with the calculation of majority of the orientation of fibres in an image, this program
169 also calculates a ‘coherency factor’ to that orientation (Lewis 2016). This factor is calculated based
170 on the amount of pixels that are in line in a particular direction and is bounded between 0 and 1;
171 with 0 and 1 indicating isotropic and anisotropic orientations, respectively (Palmieri, Lucchetti et
172 al. 2015). Twenty-five images, each of aligned and random mats were used for the analysis. All
7
173 the macroscopic images (Fig. 5 a and 5b) were taken using a Wild M3B stereomicroscope attached
175
176 Fig. 5. Macroscopic images of (a) aligned short hemp fibre mat (b) random aligned mat.
178 Aligned fibre mats (Fig. 6a) were produced using a dynamic sheet former (Fig. 6b) built by Canpa,
179 Canada. To produce fibre mats, fibre suspension (approx. 5g in 10 litres of water) was made. This
180 suspension was then pumped by the dynamic sheet former through a reciprocating nozzle onto a
181 rotating drum covered with a wire mesh which acts as a cushion for the deposited fibre (Pickering,
182 Efendy et al. 2015). The alignment of the fibres is in accordance with the nozzle and rotation of
183 the drum. A total of 45g of fibre was used for production of each mat. For the production of fibre
184 mats using DSF, well separated fibre is required to avoid blocks mainly in the flow hoses through
185 which the fibre suspension is discharged onto the rotating drum from the suspension tank.
186 Randomly aligned fibre mat was also produced by pouring fibre suspension (10g in 10 litres) onto
187 a screen with fine holes through which the water drained. The formed mat was partially drained
188 with paper towels by hand pressing. Finally, both aligned and random mats were oven dried at
190
8
191
192 Fig. 6. (a) Aligned short hemp fibre mat (b) dynamic sheet former.
194 A sheet die attached to a ThermoPrism TSE-16-TC twin-screw extruder was used to produce
195 PP/MAPP (100/7.14) sheets. In order to produce composites (except neat PP), the PP/MAPP
196 sheets and the fibre mats were weighed and arranged in stacks between two Teflon sheets before
197 inserting into a mould. The details of the mould used are available in the literature (Pickering and
198 Efendy 2016). The stacking arrangements used for the production of various composites are listed
199 in Table 1. The stacks were heated and pressed in a hot press same as that of neat PP samples (at
200 170 °C for 5 minutes at 1 MPa). Since the fibre mats are easily distorted, the production process
201 should be carried out carefully. Before the application of pressure, it was ensured that the sheets
202 were fully melted such that the matrix material consolidates sufficiently with the fibre mats.
203 Procedures detailed in ASTM D 638-03; Standard Test Method for Tensile Properties of Plastics
204 was followed for testing the specimens. An Instron-4204 tensile testing machine fitted with a 5 kN
205 load cell, operated at a constant rate of 1mm/min was used for the testing. For the measurement of
206 strain, an Instron 2630-112 extensometer with a gauge length 50 mm was attached to the central
207 part of the test specimen. Before testing, all the samples were conditioned at 23 °C ± 3 °C and
9
208 50 % ± 5% relative humidity for at least 48 hours. A total of five samples were tested from each
209 batch.
Neat PP 4 0 0 PP/PP/PP/PP
1PP*/1MAT/1PP*/
HM-15 P#-HM-15 4 3 15
1MAT/1PP*/1MAT/ 1PP*
211 Note: PP*- PP/MAPP, P#- fibres loaded perpendicular to the DSF rotation direction
214 Hemp fibres have a complex layered structure, containing primary and secondary cell walls. These
215 cell walls consist of many layers of helically wound cellulose microfibrils. The main factors that
216 determine the mechanical properties of different plant fibres are cellulose content, microfibrillar
AT/one hour 29.87 (6.7) 0.296 (0.18) 436 (236) 6.4 (2.73)
Granulated before
21.38 (4.3) 0.252(0.08) 781 (428) 12.04 (4.4)
HT
220 Table 2 displays the diameters, maximum load and mechanical properties of untreated and treated
221 fibres obtained in this work. As can be seen, the HT alkali treatment (granulated before and after
222 HT) resulted in more fibre diameter reduction compared to the AT treatment. It has been found
223 that reduction in fibre diameter is due to removal of hemicellulose and lignin (Taha, Steuernagel et
224 al. 2007, John and Anandjiwala 2008, Kabir, Wang et al. 2012, Kabir, Wang et al. 2013).
225 Although no specific studies were carried out to measure the hemicellulose or lignin content in this
226 work, it has been reported elsewhere that hemicellulose breakdown occurs easily in a high
10
227 temperature environment than at low temperature and the addition of Na2SO3 assists NaOH in the
228 removal of lignin (Beckermann 2007). It was found that HT treatment removed sufficient
229 hemicellulose and lignin from the fibres to give good fibre separation, whereas similar separation
231 From the tabulated results, it can be seen that the HT alkali treated fibre exhibited higher average
232 tensile strength and Young’s modulus compared to UT and AT alkali treated fibres, whereas the
233 average tensile strength and Young’s modulus exhibited by the AT alkali treated fibre were lower
234 than for UT fibre. This suggests that the resulting structures of the treated fibres depend on the
235 alkali treatment used. The increase in average tensile strength for HT alkali treated fibre compared
237 components) evidenced by the fibre diameter reduction after the treatment. The removal of weak
238 components from the fibre cell walls can lead to close packing of cellulose chains and possibly a
239 decrease in the microfibrillar angle. This close compaction could have enhanced the adhesions
240 between cellulose microfibrils, thereby provided better tensile properties for HT treated fibres
241 towards the loading direction compared to UT treated fibres (Efendy and Pickering 2014). As can
242 be seen in Table 2, although the diameter of AT alkali treated fibre reduced after the treatment
243 compared to UT fibre, the tensile properties of the fibre were reduced even below that of UT fibre.
244 The decrease in tensile properties associated with the fibre is thought to be due to the degradation
245 of the crystalline cellulose chains in the microfibrils or bonding between cellulose microfibrils as
246 affected by the AT treatment (Roy, Chakraborty et al. 2012, Efendy and Pickering 2014).
247 A Student’s t-test was carried out for comparing HT alkali treated fibres with different granulation
248 sequence and it was found that the sequence of granulation had no significant effect on tensile
249 properties of the fibres. Weibull modulus, Weibull characteristic strength and experimental
250 average tensile strength are displayed in Table 3. As expected, the characteristic strength has the
251 same trend as that of average tensile strength. The Weibull modulus for the fibres varied from 1.47
252 to 2.05. These values are comparable with those reported in the literature for cellulosic fibres
254 Table 3: Comparison of Weibull parameters with experimental tensile strength for hemp samples
Characteristic Strength Average Tensile
Hemp Samples Weibull Modulus, w
(MPa), σ0 Strength (MPa)
11
UT 1.68 576 517
256
257 Fig. 7. X-ray diffraction curves for untreated, atmospheric temperature treated and high
258 temperature treated fibres.
259 Crystallinity index indicates the degree of crystallinity (Mwaikambo and Ansell 2002, Ouajai and
260 Shanks 2005). The X-ray diffraction profiles (curves) of UT, AT/one hour, AT/two hours and
261 granulated after HT treated hemp fibres are shown in Fig. 7. The Ic values were calculated from
262 maximum and minimum intensity crystallographic peaks for each profile which are around 2θ =
UT 64.87
265
266 As can be seen in Table 4, alkali treatments improved the Ic values except for AT/two hours. The
267 higher Ic value of HT alkali treated fibre compared to the UT fibre would be expected due to the
268 removal of non-crystalline materials and possibly better packing of cellulose chains within the
12
269 fibre (Beckermann 2007). Although the Ic value of AT/one hour alkali treated fibres was higher
270 compared to UT fibre, as discussed earlier the tensile strength of this fibre was lower than that of
271 UT fibre. This suggests that chain scission could have overridden the influence of increased
272 crystallinity (Islam, Pickering et al. 2010, Islam, Pickering et al. 2011, Sawpan, Pickering et al.
273 2011, Ghazali and Efendy 2016). It has been reported elsewhere that the degradation rate of
274 cellulose in alkali is influenced by fibrillar morphology; a more ordered physical structure impedes
275 degradation (Knill and Kennedy 2003). This supports the production of better-packed cellulose
276 chains with HT treatment which would have impeded the diffusion of alkali reducing cellulose
279 Peaks (Fig. 8) in the regions 1730-1740 cm-1 and 1200-1300cm-1 indicate the hemicellulose and
280 lignin components through the presence of C=O linkages (Abraham, Deepa et al. 2011, Chen, Yu
281 et al. 2011). Peaks at 1737 cm-1, 1252 cm-1 and 1201 cm-1 for the UT fibres, became smaller for
282 the AT alkali treated fibres and were not visible for HT alkali treated fibres. Reduction of peak
283 heights supports that alkali treatment removed hemicellulose and lignin, with more removal
284 occurring in HT alkali treated fibres compared to AT alkali treated fibres (Olsson and Salmén
285 2004, Li and Pickering 2008, Peng, Ren et al. 2009, Abraham, Deepa et al. 2011, Islam, Pickering
286 et al. 2011, Kabir, Wang et al. 2013). Similarly, the smaller peaks in the range between 1280 cm-1
287 and 1330 cm-1 for HT alkali treated fibre compared to AT alkali treated and UT fibres further
288 support that HT treatment removes more hemicellulose than AT alkali treated fibres (Taha,
289 Steuernagel et al. 2007). The intensity of peaks between 1630 cm-1 and 1650 cm-1 slightly
290 increased after alkali treatment, which may be due to water molecules formed by the reactions
291 between sodium hydroxide and cellulosic hydroxyl groups (Le Troedec, Sedan et al. 2008).
13
292
293 Fig. 8. The FTIR spectra of untreated, AT/one hour treated, and granulated after HT treated fibres.
295 Typically, for NPFs, there are three main stages of degradation where most of the weight loss
296 occurs: 50-100 °C due to evaporation of moisture in the fibres, 200-350 °C due to hemicellulose
297 decomposition and 300-500 °C mainly due to degradation of lignin and cellulose (Sun, Tomkinson
298 et al. 2000, Dahiya and Rana 2004, Methacanon, Weerawatsophon et al. 2010).
299
300 Fig. 9. TGA thermograms for UT, AT/one hour treated, and granulated after HT treated fibres.
14
301
302 Fig. 10. Weight loss summary for different samples.
303 Figures 9 and 10 (Fig. 9 and Fig. 10) show the TGA thermograms and weight loss summary for
304 UT fibres, AT and HT alkali treated fibres. From Fig. 10, it can be seen that the initial 10 %
305 weight loss occurred only at 297 °C and 329 °C for the AT and HT alkali treated fibres
306 respectively compared to 288 °C for UT fibres supporting the overall improved thermal stability of
307 the fibres. The improved thermal stability of the fibres is likely to be due to thermally unstable
308 components (hemicellulose and pectin) being removed from the fibres due to alkali treatment
309 (Beckermann 2007), with more removal occurring with HT alkali treatment compared to AT
310 treated fibres as supported by FTIR analysis. As the temperature further increased above 360 °C,
311 the weight loss was lower for UT fibres compared to treated fibres, which may be due to a stable
312 lignocellulose complex formed at higher temperatures that prevented this lignin-rich fibre from
313 further weight loss above 360 °C (Islam 2008). Also, at higher temperatures, it was found that the
314 weight percentage loss was higher for HT alkali treated fibres compared to AT alkali treated fibres
315 (Fig. 10). This higher weight percentage loss for HT alkali treated fibres above 360 °C supports
316 that the greater removal of lignin from the fibre was by HT alkali treatment compared to AT alkali
317 treatment (Beckermann 2007, Ghazali and Efendy 2016). This greater removal was further
318 supported by the higher amount of residue obtained after the TGA analysis for AT treated fibres
15
320 3.5 SEM microscopy of hemp fibre
321
322 Fig. 11. SEM images of hemp fibre surfaces (a) UT fibre (b) AT/one hour alkali treated fibre (c)
323 granulated before HT alkali treated fibre (d) granulated after HT alkali treated fibre.
324 Fig. 11 shows the SEM micrographs of hemp fibre. As can be seen in Fig. 11a, the UT fibres are
325 mostly bundle form; substances known to include lignin, pectin, hemicellulose and other non-
326 strengthening components are localised on their surfaces (Ghazali and Efendy 2016). Alkali
327 treated hemp fibres appeared to have undergone some degree of fibre separation known to occur
328 due to the removal of some of these components (Fig. 11b, 11c and 11d). However, it was found
329 that the AT treatment resulted in very little separation of fibres compared to the HT alkali
330 treatment (Fig. 11b). It was also found that the fibres granulated after HT alkali treatment were
331 better separated compared to the fibres granulated before HT alkali treatment, which is evident
334 It was found that only fibres granulated after HT alkali treatment were sufficiently separated to
335 successfully be used in a dynamic sheet former to form short hemp fibre mats (Fig. 5a and Fig.
16
336 6a). Fig. 12 shows the fibre orientation distribution profiles obtained for the mats analysed using
337 OrientationJ (ImageJ). As can be seen for the random mats, there appeared only a relatively small
338 broad peak (almost a flat curve). In contrast, there appeared a sharp predominant peak for the mats
339 produced using DSF around 0 ° (± 5 °), i.e. the preferred orientation direction. The coherency
340 factors generated by the OrientationJ program for the DSF mats and random mats were 0.23 (±
341 0.028) and 0.11 (0.038), respectively. The predominant peak and higher coherency factor for the
342 DSF mats compared to the broad peak and lower coherency factor for the random mats support the
343 potential of dynamic sheet former to produce aligned short hemp fibre mats.
344
347
348 Fig. 13. Typical stress-strain curves for polypropylene reinforced with 15 wt% hemp fibre loaded
349 parallel and perpendicular to the DSF rotation (main fibre orientation) direction.
17
350 Composites were tested parallel and perpendicular to the main fibre orientation direction (DSF
351 rotation direction). Fig. 13 shows the stress versus strain curves for the composites with fibre
352 content of 15 wt% along with that of neat PP for comparison purposes. Neat PP extended in a
353 ductile manner to high strain without fail, whereas the incorporation of fibres caused PP to fail in
a b
355
356 Fig. 14. Graph representing (a) tensile strength and (b) Young’s modulus for the hemp composites
357 tested parallel and perpendicular to the DSF rotation direction.
358 Tensile properties for neat PP along with composites are presented in Fig. 14. The maximum
359 tensile strength and Young’s modulus of the composite were approximately 59 and 120 % higher
360 than for neat PP. From the results, it can be seen that composite tested parallel to the main fibre
361 orientation direction exhibited higher tensile properties compared to the composite tested
362 perpendicular to the main fibre orientation direction. This further supports the potential of DSF to
363 produce aligned fibre mats. It has been reported that the best mechanical properties for composites
364 are generally being obtained when the fibres are aligned parallel to the direction of the applied
365 load (Herrera-Franco and Valadez-Gonzalez 2005, Pickering, Efendy et al. 2015).
366 4. Conclusion
367 The high temperature treatment at 120 °C using 5wt% NaOH and 2wt% Na 2SO3, with a fibre to
368 solution ratio of 1:8, improved the tensile strength and Young’s modulus of hemp fibres by 51%
369 and 62% respectively compared to untreated fibre. In contrast, tensile strength and Young’s
18
370 modulus of ambient temperature treated hemp fibre were lower than that of untreated fibres by
371 16% and 14% respectively. SEM, XRD, FTIR and TGA analyses support that the high
372 temperature treatment removes more non-strengthening components from the fibres compared to
373 ambient temperature treatment. Improvement of fibre strength with high temperature alkali
374 treatment compared with the reduction of fibre strength obtained with ambient temperature alkali
375 treatment suggests better packing of cellulose chains occur for high temperature treatment
376 providing better resistance to cellulose degradation. Therefore, high temperature treatment is
377 recommended for producing strong and stiff fibres for use in natural plant fibre composites.
378 Overall, significantly aligned short hemp fibre mats from high strength hemp fibres was produced
380 Acknowledgements
381 This research received no specific grant from any agency in the public, commercial, or not-for-
382 profit sectors. However, the authors would like to thank to the University of Waikato’s
384 References
385 Abraham, E., B. Deepa, L. Pothan, M. Jacob, S. Thomas, U. Cvelbar and R. Anandjiwala (2011).
386 "Extraction of nanocellulose fibrils from lignocellulosic fibres: a novel approach." Carbohydrate
387 Polymers 86(4): 1468-1475.
388 Abràmoff, M. D., P. J. Magalhães and S. J. Ram (2004). "Image processing with ImageJ."
389 Biophotonics international 11(7): 36-42.
390 ASTM, D. (1986). "3379-75." Standard Test Method for Tensile Strength and Young's Modulus
391 for High-Modulus Single-Filament Materials, Annual Book of ASTM Standards,(May 1989) 8:
392 128-131.
393 Beckermann, G. (2007). "Performance of hemp-fibre reinforced polypropylene composite
394 materials."
395 Beckermann, G. and K. L. Pickering (2008). "Engineering and evaluation of hemp fibre reinforced
396 polypropylene composites: fibre treatment and matrix modification." Composites Part A: Applied
397 Science and Manufacturing 39(6): 979-988.
398 Beg, M. D. H. (2007). The improvement of interfacial bonding, weathering and recycling of wood
399 fibre reinforced polypropylene composites, The University of Waikato.
400 Biagiotti, J., D. Puglia, L. Torre, J. M. Kenny, A. Arbelaiz, G. Cantero, C. Marieta, R. Llano‐Ponte
401 and I. Mondragon (2004). "A systematic investigation on the influence of the chemical treatment
402 of natural fibers on the properties of their polymer matrix composites." Polymer Composites 25(5):
403 470-479.
404 Bledzki, A. and J. Gassan (1999). "Composites reinforced with cellulose based fibres." Progress in
405 polymer science 24(2): 221-274.
406 Chen, W., H. Yu, Y. Liu, P. Chen, M. Zhang and Y. Hai (2011). "Individualization of cellulose
407 nanofibers from wood using high-intensity ultrasonication combined with chemical
408 pretreatments." Carbohydrate Polymers 83(4): 1804-1811.
409 Dahiya, J. and S. Rana (2004). "Thermal degradation and morphological studies on cotton
410 cellulose modified with various arylphosphorodichloridites." Polymer International 53(7): 995-
411 1002.
412 Efendy, M. A. and K. Pickering (2014). "Comparison of harakeke with hemp fibre as a potential
413 reinforcement in composites." Composites Part A: Applied Science and Manufacturing 67: 259-
414 267.
19
415 French, A. D. (2014). "Idealized powder diffraction patterns for cellulose polymorphs." Cellulose
416 21(2): 885-896.
417 Gesellchen, F., A. Bernassau, T. Dejardin, D. Cumming and M. Riehle (2014). "Cell patterning
418 with a heptagon acoustic tweezer–application in neurite guidance." Lab on a Chip 14(13): 2266-
419 2275.
420 Ghazali, M. and A. Efendy (2016). Bio-composites materials from engineered natural fibres for
421 structural applications, University of Waikato.
422 Herrera-Franco, P. and A. Valadez-Gonzalez (2005). "A study of the mechanical properties of
423 short natural-fiber reinforced composites." Composites Part B: Engineering 36(8): 597-608.
424 Islam, M., K. Pickering and N. Foreman (2010). "Influence of alkali treatment on the interfacial
425 and physico-mechanical properties of industrial hemp fibre reinforced polylactic acid composites."
426 Composites Part A: Applied Science and Manufacturing 41(5): 596-603.
427 Islam, M. S. (2008). The influence of fibre processing and treatments on hemp fibre/epoxy and
428 hemp fibre/PLA composites, the University of Waikato.
429 Islam, M. S., K. L. Pickering and N. J. Foreman (2011). "Influence of alkali fiber treatment and
430 fiber processing on the mechanical properties of hemp/epoxy composites." Journal of Applied
431 Polymer Science 119(6): 3696-3707.
432 John, M. J. and R. D. Anandjiwala (2008). "Recent developments in chemical modification and
433 characterization of natural fiber‐reinforced composites." Polymer composites 29(2): 187-207.
434 Kabir, M., H. Wang, K. Lau and F. Cardona (2012). "Chemical treatments on plant-based natural
435 fibre reinforced polymer composites: An overview." Composites Part B: Engineering 43(7): 2883-
436 2892.
437 Kabir, M., H. Wang, K. Lau and F. Cardona (2013). "Effects of chemical treatments on hemp fibre
438 structure." Applied Surface Science 276: 13-23.
439 Kabir, M., H. Wang, K. Lau and F. Cardona (2013). "Tensile properties of chemically treated
440 hemp fibres as reinforcement for composites." Composites Part B: Engineering 53: 362-368.
441 Kabir, M., H. Wang, K. Lau, F. Cardona and T. Aravinthan (2012). "Mechanical properties of
442 chemically-treated hemp fibre reinforced sandwich composites." Composites Part B: Engineering
443 43(2): 159-169.
444 Knill, C. J. and J. F. Kennedy (2003). "Degradation of cellulose under alkaline conditions."
445 Carbohydrate Polymers 51(3): 281-300.
446 Le, T. M. (2016). Harakeke fibre as reinforcement in epoxy matrix composites and its
447 hybridisation with hemp fibre, University of Waikato.
448 Le, T. M. and K. L. Pickering (2015). "The potential of harakeke fibre as reinforcement in polymer
449 matrix composites including modelling of long harakeke fibre composite strength." Composites
450 Part A: Applied Science and Manufacturing 76: 44-53.
451 Le Troedec, M., D. Sedan, C. Peyratout, J. P. Bonnet, A. Smith, R. Guinebretiere, V. Gloaguen
452 and P. Krausz (2008). "Influence of various chemical treatments on the composition and structure
453 of hemp fibres." Composites Part A: Applied Science and Manufacturing 39(3): 514-522.
454 Lewis, D. D. J. (2016). Interlaminar reinforcement of carbon fiber composites from unidirectional
455 prepreg utilizing aligned carbon nanotubes, Massachusetts Institute of Technology.
456 Li, X., L. G. Tabil and S. Panigrahi (2007). "Chemical treatments of natural fiber for use in natural
457 fiber-reinforced composites: a review." Journal of Polymers and the Environment 15(1): 25-33.
458 Li, Y. and K. L. Pickering (2008). "Hemp fibre reinforced composites using chelator and enzyme
459 treatments." Composites Science and Technology 68(15): 3293-3298.
460 Methacanon, P., U. Weerawatsophon, N. Sumransin, C. Prahsarn and D. Bergado (2010).
461 "Properties and potential application of the selected natural fibers as limited life geotextiles."
462 Carbohydrate Polymers 82(4): 1090-1096.
463 Mwaikambo, L. Y. and M. P. Ansell (2002). "Chemical modification of hemp, sisal, jute, and
464 kapok fibers by alkalization." Journal of applied polymer science 84(12): 2222-2234.
465 Olsson, A.-M. and L. Salmén (2004). "The association of water to cellulose and hemicellulose in
466 paper examined by FTIR spectroscopy." Carbohydrate research 339(4): 813-818.
467 Ouajai, S. and R. Shanks (2005). "Composition, structure and thermal degradation of hemp
468 cellulose after chemical treatments." Polymer degradation and stability 89(2): 327-335.
469 Oushabi, A., S. Sair, F. O. Hassani, Y. Abboud, O. Tanane and A. El Bouari (2017). "The effect of
470 alkali treatment on mechanical, morphological and thermal properties of date palm fibers (DPFs):
471 Study of the interface of DPF–Polyurethane composite." South African Journal of Chemical
472 Engineering 23: 116-123.
473 Palmieri, V., D. Lucchetti, A. Maiorana, M. Papi, G. Maulucci, F. Calapà, G. Ciasca, R. Giordano,
474 A. Sgambato and M. De Spirito (2015). "Mechanical and structural comparison between primary
475 tumor and lymph node metastasis cells in colorectal cancer." Soft matter 11(28): 5719-5726.
20
476 Peng, F., J.-L. Ren, F. Xu, J. Bian, P. Peng and R.-C. Sun (2009). "Comparative study of
477 hemicelluloses obtained by graded ethanol precipitation from sugarcane bagasse." Journal of
478 agricultural and food chemistry 57(14): 6305-6317.
479 Pickering, K. (2008). Properties and performance of natural-fibre composites, Elsevier.
480 Pickering, K., G. Beckermann, S. Alam and N. Foreman (2007). "Optimising industrial hemp fibre
481 for composites." Composites Part A: Applied Science and Manufacturing 38(2): 461-468.
482 Pickering, K. and M. A. Efendy (2016). "Preparation and mechanical properties of novel bio-
483 composite made of dynamically sheet formed discontinuous harakeke and hemp fibre mat
484 reinforced PLA composites for structural applications." Industrial Crops and Products 84: 139-
485 150.
486 Pickering, K., M. A. Efendy and T. Le (2015). "A review of recent developments in natural fibre
487 composites and their mechanical performance." Composites Part A: Applied Science and
488 Manufacturing.
489 Pickering, K. and T. M. Le (2016). "High performance aligned short natural fibre–Epoxy
490 composites." Composites Part B: Engineering 85: 123-129.
491 Püspöki, Z., M. Storath, D. Sage and M. Unser (2016). Transforms and operators for directional
492 bioimage analysis: a survey. Focus on Bio-Image Informatics, Springer: 69-93.
493 Rezakhaniha, R., A. Agianniotis, J. T. C. Schrauwen, A. Griffa, D. Sage, C. v. Bouten, F. Van de
494 Vosse, M. Unser and N. Stergiopulos (2012). "Experimental investigation of collagen waviness
495 and orientation in the arterial adventitia using confocal laser scanning microscopy." Biomechanics
496 and modeling in mechanobiology 11(3-4): 461-473.
497 Roy, A., S. Chakraborty, S. P. Kundu, R. K. Basak, S. B. Majumder and B. Adhikari (2012).
498 "Improvement in mechanical properties of jute fibres through mild alkali treatment as
499 demonstrated by utilisation of the Weibull distribution model." Bioresource technology 107: 222-
500 228.
501 Sawpan, M. A., K. L. Pickering and A. Fernyhough (2011). "Effect of various chemical treatments
502 on the fibre structure and tensile properties of industrial hemp fibres." Composites Part A: Applied
503 Science and Manufacturing 42(8): 888-895.
504 Schneider, C. A., W. S. Rasband and K. W. Eliceiri (2012). "NIH Image to ImageJ: 25 years of
505 image analysis." Nature methods 9(7): 671-675.
506 Segal, L., J. Creely, A. Martin Jr and C. Conrad (1959). "An empirical method for estimating the
507 degree of crystallinity of native cellulose using the X-ray diffractometer." Textile research journal
508 29(10): 786-794.
509 Shah, D. U., F. Vollrath, J. Stires and D. D. Deheyn (2015). "The biocomposite tube of a
510 chaetopterid marine worm constructed with highly-controlled orientation of nanofilaments."
511 Materials Science and Engineering: C 48: 408-415.
512 Sun, R., J. Tomkinson and G. L. Jones (2000). "Fractional characterization of ash-AQ lignin by
513 successive extraction with organic solvents from oil palm EFB fibre." Polymer Degradation and
514 Stability 68(1): 111-119.
515 Sunny, T., K. L. Pickering and S. H. Lim (2017). Alignment of Short Fibres: An Overview.
516 Processing and Fabrication of Advanced Materials-XXV.
517 Sunny, T., K. L. Pickering and S. H. Lim (2018). Fibre Orientation Distribution Assessment of
518 Dynamically Sheet Formed Hemp Fibre Mats by Image Analysis. Applied Mechanics and
519 Materials, Trans Tech Publ.
520 Taha, I., L. Steuernagel and G. Ziegmann (2007). "Optimization of the alkali treatment process of
521 date palm fibres for polymeric composites." Composite Interfaces 14(7-9): 669-684.
522 Zafeiropoulos, N. and C. Baillie (2007). "A study of the effect of surface treatments on the tensile
523 strength of flax fibres: Part II. Application of Weibull statistics." Composites Part A: Applied
524 Science and Manufacturing 38(2): 629-638.
525
526
21