Download as pdf or txt
Download as pdf or txt
You are on page 1of 27

Version of Record: https://www.sciencedirect.

com/science/article/pii/S0926669020301011
Manuscript_e167e17eee96403e38bc057ac43541e2

1 Biochar filled high-density polyethylene composites with excellent properties:


2 Towards maximizing the utilization of agricultural wastes
3
4 Qingfa Zhanga,b, Donghong Zhanga, Hang Xua, Wenyu Lua, Xiajin Rena, Hongzhen Caia*,
5 Hanwu Leib**, Erguang Huob, Yunfeng Zhaob, Moriko Qianb, Xiaona Lina,b, Elmar M. Villotab,
6 Wendy Mateob
7
aSchool
8 of Agricultural Engineering and Food Science, Shandong Research Center of Engineering & Technology
9 for Clean Energy, Shandong University of Technology, Zibo 255000, China
bDepartment
10 of Biological Systems Engineering, Washington State University, Richland, WA,99354, USA;
11 ∗ Corresponding author: Email address: chzh666666@126.com (H.Cai)
12 **Corresponding author. Tel.: +1 509 372 7628; fax +1 509 372 7690. E-mail: hlei@wsu.edu (H. Lei)
13
14 Abstract: Biochar derived from agricultural wastes was used to reinforce high-density
15 polyethylene (HDPE) to obtain composites abbreviated as B10, B20, B30, B40, B50, B60, and
16 B70. Investigating the mechanical, thermal, water absorption and flame retardant properties of the
17 composites is one of the objectives while maximizing the utilization of agricultural wastes is the
18 ultimate goal of this work. It was found that flexural properties, tensile properties, storage
19 modulus, elasticity, creep resistance and anti-stress relaxation ability of HDPE were improved by
20 the inclusion of biochar, and excellent mechanical properties were obtained in 50% even 60%
21 biochar added composites because of good dispersion and unique structure shown in the interface.
22 The composites achieved good flexural strength of 34.95 MPa in B50, flexural modulus of 1.79
23 GPa in B40, tensile strength of 29.05 MPa in B40, and tensile modulus of 2.03 GPa. Additionally,
24 the thermal and flame retardant properties (limited oxygen index of 25.06% in B70) increased for
25 the biochar added composites as the biochar loading increased due to the high thermal stability of
26 biochar, although biochar had a negative effect on water-resistance of the composites. The results
27 revealed that the ultimate goal was achieved in terms of producing composites with excellent
28 mechanical, thermal, water absorption and flame retardant properties while maximizing the
29 utilization of agricultural wastes as a rational balance.
30
31 Keywords: Agricultural wastes; maximization; biochar; composites; properties evaluation
32
33 1. Introduction
34 The demand for fossil fuels increases dramatically along with worldwide population growth,
35 economic progress, and technological advancement. One great challenge is the pollution caused
36 by the fossil fuel with its depletion becoming another problem that cannot be ignored (Das et al.,
37 2015). On the other hand, the urgent demand for food due to population growth has contributed to
38 the agglomeration of a large number of agricultural wastes. In traditional agricultural production,
39 incineration and discarding are the main ways in disposing of agricultural wastes (Demirbas,
40 2011). Nevertheless, these two ways pollute the air and water and aggravate the greenhouse effect
41 showing a huge threat to agricultural product security and environmental safety (Lim et al., 2016).
42 Hence, the effective disposal of agricultural wastes is an urgent need. In fact, agricultural wastes
43 have great potential for resource utilization due to their rich organic matters including cellulose,
44 hemicellulose, lignin, etc. Far more important is that agricultural wastes are renewable comparing

© 2020 published by Elsevier. This manuscript is made available under the Elsevier user license
https://www.elsevier.com/open-access/userlicense/1.0/
45 with fossil fuels. In a nutshell, incineration and discarding of agricultural wastes are the waste of
46 resources. On the other hand, polyolefine plastics like polyethylene (PE), polypropylene (PP),
47 polyvinyl chloride (PVC), etc. have become an irreplaceable raw material for industrial production
48 and daily life due to their excellent characteristics with technological advancement (Chow et al.,
49 2018). Be that as it may, polyolefine plastics show non-degradability which is responsible for
50 serious environmental pollution. Likewise, the main disposals of waste plastics, such as
51 incineration and landfilling, have created more serious problems of polluting the air and water
52 seriously (Fa et al., 2019). The recycling of waste plastics has become another great problem
53 affecting environmental safety. Given all that, the effective and practical disposal of agricultural
54 and plastic wastes is a real challenge especially with the current pressure of efforts on
55 environmental protection.
56 In the view of efficiently utilizing plastics and agricultural wastes, biocomposites made great
57 progress worldwide, especially in North America over the past decades. Normally, some
58 biocomposites are prepared with plastic wastes like PE, PP, PVC, etc. and agricultural wastes like
59 corn stalk, rice husk, peanut shell, etc. by extrusion, hot pressing and injection molding (Arrigo et
60 al., 2019; Ferreira et al., 2019; Zhang et al., 2019b). The biocomposites have been used in
61 construction, transportation, industry, and decoration due to their corrosion resistance, easy
62 processing, and acid and alkali resistance. Importantly, the biocomposites can be recycled and
63 reused discharging no pollution in the production process (Wang et al., 2019). Nevertheless,
64 because of the incompatibility between agricultural waste fibers and plastics, the interfacial
65 adhesion of these biocomposites is typically weak causing the poor mechanical properties
66 subsequently (Turku et al., 2017). To overcome this issue, interfacial compatibilizers such as
67 maleic anhydride are added to the biocomposites and the mechanical properties are efficiently
68 improved. Bajwa et al. studied the effect of maleic anhydride grafting polyethylene (MAPE) on
69 the properties of high-density polyethylene (HDPE)/wood fiber composites, and better mechanical
70 properties were obtained with the addition of MAPE (Bajwa et al., 2019). Nevertheless, the
71 addition of costly interfacial compatibilizers increased the production cost of biocomposites
72 indicating a new problem. The balance of properties and the economic efficiency of biocomposites
73 has become a new challenge. Therefore, lots of studies have been done to solve this problem.
74 Among them, the utilization of biochar for biocomposites attracted extensive attention due to its
75 porous structure and large surface area in recent years (Das et al., 2015; Nan et al., 2016).
76 All the agricultural wastes could be converted to biochar by a thermochemical process known
77 as carbonization or pyrolysis with abundance feedstock sources, which lower the production cost
78 of biochar. Moreover, biochar reinforced biocomposites achieve the balance between properties
79 and cost (Behazin et al., 2017; Das et al., 2019). On the one hand, a physical/mechanical
80 interlocking structure (Das et al., 2017) formed by porous biochar in plastics effectively transfers
81 stress and improves the mechanical properties of biocomposites. Bajwa et al. and Das et al. found
82 that the coupling agent was not effective in biochar added composites indicating unnecessary of
83 using interfacial compatibilizers which reduced the cost (Bajwa et al., 2019; Das et al., 2016b).
84 Subsequently, lots of efforts have been done in biochar biocomposites in terms of the unique
85 advantages of biochar. The study of Giorcelli et al. showed that biochar was a cheap and
86 environmental friendly filler to improve polymer mechanical properties (Giorcelli et al., 2019). Ho
87 et al. used bamboo biochar to reinforce polylactic acid (PLA), and an excellent improvement in
88 the properties of PLA was obtained (Ho et al., 2015). Das et al. thought that the utilization of

2
89 biochar to prepare biocomposites was a sustainable and resilient approach to manage agricultural
90 wastes (Das et al., 2015). The development of biochar biocomposites attracted more and more
91 scholars to participate in the study, and the number of available documents has been more than 50
92 in the last five years. The biochar sources (Poulose et al., 2018; Santhiago et al., 2018; Zhang et al.,
93 2018) are not limited to agricultural wastes involving forestry wastes, grasses, manure, even bones,
94 etc. Recently, Mohanty et al. published a paper in Science demonstrating that biochar was a new
95 sustainable and functional filler material in biocomposites (Mohanty et al., 2018). Besides, the
96 effects of biochar loading, biochar feedstock sources and types of polymers on biocomposites are
97 also studied. Though researchers have a universal agreement that biochar loading has the greatest
98 influence on biochar biocomposites, the biochar loading of most investigations is low. The
99 maximum biochar loadings were 35% (Das et al., 2016a), 10% (Ho et al., 2015), 20% (Mashouf
100 Roudsari et al., 2017) shown by available documents respectively. In view of production cost of
101 polymers, it is necessary to increase the biochar loading in biocomposites. High biochar loadings
102 help maximize the utilization of agricultural wastes on the premise of ensuring environmental
103 safety and reducing costs. To date, the effect of high biochar loadings on biochar biocomposites is
104 rarely investigated. Hence, the focus of this work is to optimize biocomposites by regulating
105 biochar loadings. A balance of biocomposites properties and economic efficiency is expected to
106 achieve in terms of maximizing the use of agricultural wastes while minimizing the production
107 cost.
108
109 2. Materials and methods
110 2.1 The constituents
111 Rice husk was used as feedstock to produce biochar at 600 °C using a miniature box furnace
112 (KSL-1100X-S). To get a higher yield, the pyrolysis temperature was increased from 25 °C
113 (ambient temperature) to 600 °C at a low heating rate (5 °C/min) under 20 mL/min nitrogen
114 atmosphere and held for 2 h to obtain the biochar sample. The biochar sample was sieved over 150
115 μm screen to keep biochar particle size below150 μm. The main parameters of the biochar sample
116 and ashes were shown in Table S1. HDPE (9001) pellets with a melt index of 0.05 g/10min and a
117 density of 0.950 g/cm3 were purchased from USI Corporation. The biochar and HDPE were dried
118 at 105 °C and 50 °C for 24 h prior to compounding with each other.
119
120 2.2 Composites processing
121 Dried biochar particles and HDPE pellets were put into a high-speed mixer for 10 min to get
122 the uniform blends. Subsequently, a micro twin-screw extruder (WLG10G) was used to melt and
123 mix the composites blends, and then a micro-injection molding machine (WZS10D) was used to
124 get cuboid shaped and dog bone shaped composite samples. The extruding zone and blending
125 zone of the extruder were run at 30 rpm and 180 °C, the four temperature zones of injection
126 moulding were all set to 180 °C and the mold was kept at 50 °C. To maximize the utilization of
127 agricultural wastes, increasing the loading of biochar as much as possible is necessary for this
128 study. It was found that 70 wt% loading of biochar was the maximum load that can be afforded in
129 the composites in view of molding and deformation as shown in Figure 1. Based on this, the
130 biochar/HDPE mixtures contained biochar ranging from 10 to 70 wt% with an interval of 10 wt%
131 each, and the composite samples under different biochar loadings were named as B10, B20, B30,
132 B40, B50, B60, and B70 respectively.

3
133
134 2.3 Composites characterization
135 Chemical characterization of HDPE and different biocomposites was conducted using
136 Fourier transform infrared spectroscopy (FTIR) and X-Ray diffraction (XRD). FTIR of the
137 samples was performed in a Fourier transform infrared spectroscopy (Nicolet 5700, Thermo
138 Fisher Nicolet, USA) with a scanning range from 4000 cm-1 to 400 cm-1. XRD of the samples was
139 performed in a polycrystalline X-ray diffractometer (Bruker AXS D8 Advance, Germany) using
140 CuKα radiation with 2θ varying between 5° and 60° at 5 °/min. Microstructure characterization of
141 fractured surfaces of different biocomposites was conducted in a scanning electron microscope
142 (SEM) (FEI Sirion 200, USA) with a scanning voltage of 3.0 kV.
143 Thermal characterization of HDPE and different biocomposites was conducted using
144 Differential scanning calorimetry (DSC) and thermogravimetric analysis (TGA). DSC of the
145 samples was performed in a differential scanning calorimeter analyzer (DSC-Q100, TA Instrument,
146 USA), and all the samples were heated to 180 °C and then cooled to 100 °C and then heated to
147 180 °C, all at 5 °C/min. TGA of the samples was performed in a synchronous thermal analyzer
148 (STA 449, NETZSCH), and all the samples were heated from 30 to 600 °C under the 20 mL/min
149 nitrogen atmosphere with a heating rate of 10 °C/min.
150 Mechanical characterization of HDPE and different biocomposites was conducted by
151 dynamic mechanical analysis (DMA) and static mechanical analysis (SMA). DMA including
152 viscoelastic behavior, creep compliance and relaxation modulus was performed using a DMA Q
153 800 (TA instrument, USA). The samples were heated from -50 to 150 °C to obtain the viscoelastic
154 behavior in a single cantilever mode with the heating rate of 5 °C/min, and the 30 min creep
155 compliance and relaxation modulus of the samples were obtained using the Creep TTS and Stress
156 Relaxation TTS models of DMA under 30 °C. SMA including flexural and tensile properties was
157 performed using an electronic universal testing machine (WDW1020, Changchun Kexin Co., Ltd.,
158 China). Five specimens of each composite sample were tested. The average values were reported
159 with their standard errors, and the results were analyzed with SPSS using a one-way analysis of
160 variance (ANOVA). The statistical analysis showed that the composites with varied loadings of
161 biochar had a significant difference in flexural and tensile properties (p<0.05).
162 HDPE and different biocomposites were completely immersed in water at ambient
163 temperature for durations of up to 4 weeks, and the samples were removed and wiped to dry to
164 measure the weights using a precision weigh balance. Statistical analysis showed that the
165 composites with varied biochar loadings had a significant difference in water absorption (p<0.05).
166 The water uptake of samples was calculated by:
167 Water Absorption (%) = (W1 - W0) / W0 × 100%
168 where W1 and W0 are the weights of before and after soaking respectively.
169
170 The flame retardancy of HDPE and different biocomposites was characterized by a limited
171 oxygen index using an oxygen index tester (ZR-01, China), and the flow rate of oxygen and
172 oxygen-nitrogen mixture were 3 L/min and 20 L/min, respectively. Statistical analysis showed that
173 the composites with varied biochar loadings had a significant difference in the limited oxygen
174 index (p<0.05). The limited oxygen index of samples was calculated by:
175 Limited Oxygen Index (%) = [O2] / ([O2] + [N2]) × 100%
176 where [O2] and [N2] are the flow rate of oxygen and nitrogen, respectively

4
177
178 3. Results and discussion
179 3.1 XRD
180 Figure 2 provides the results obtained from the XRD analysis of HDPE and composite
181 samples. From the figure, one can see two typical characteristic peaks at 2θ of 21.6° and 24°
182 which corresponded to (110) and (200) crystal planes. Many documents confirmed that these two
183 typical crystal planes belonged to neat PE (Arnandha et al., 2017). It should be noted that the
184 position of the two typical characteristic peaks was not changed with the addition of loading
185 varied biochars indicating that the two major peaks were both contributed by HDPE and the
186 addition of biochar did not change the crystal planes of HDPE. A big intensity drop was observed
187 in B10, which was due to that biochar addition even at low loading (10%) reduced the crystallinity
188 degree (Xc) of HDPE as shown in Table 1. Additionally, the intensity of these two peaks in
189 composites was reduced as the loading of biochar increased. Since the biochar is amorphous in
190 nature (Das et al., 2015), the increment of amorphous biochar and reduction of Xc with the
191 increase of biochar loadings should be responsible for the reduction of major peaks in composites
192 (Zhang et al., 2019a).
193
194 3.2 FTIR
195 Figure 3 illustrates the FTIR spectra of HDPE and composite samples. As can be observed
196 from Figure 3 that three typical peaks at 2920, 2850 and 1470cm-1 were shown in the FTIR spectra
197 of HDPE and these three peaks can be assigned to the C-H stretching vibration of aliphatic
198 structures (Hahn et al., 2019). Highly biochar added composites, especially B10, showed weaker
199 peaks at 2920, 2850 and 1470cm-1, which was attributed to the lack of HDPE in composites.
200 Interestingly, peaks involving asymmetrical O-H stretching vibration and C-O stretching vibration
201 around 3400cm-1 and 1080cm-1 were weakly detected in composites samples (Song et al., 2019). In
202 addition, both two peaks in the composites were contributed by the biochar. The asymmetrical
203 O-H stretching vibration was from phenolic hydroxyl and alcohol hydroxyl in rice husk, and C-O
204 stretching vibration was caused by carbohydrates. The weakening of these two functional groups
205 indicates that the carbonization at 600 °C reduced the polarity of rice husk and resulted in a low
206 hydrophilicity. The similar results were also obtained in Figure S2. The contact angle analysis
207 showed that biochar added composites presented lower hydrophilicity compared to rice
208 husk/HDPE composites, which indicates that biochar added composites had better mechanical
209 properties than rice husk/HDPE composites. The document demonstrated that the coupling agent
210 showed little effect in biochar/HDPE composites, which confirmed the analysis of this study
211 (Bajwa et al., 2019). Moreover, changes in the peak intensities, shifts and appearances or
212 disappearances were not observed among the samples, which was evident that no chemical
213 reaction occurred between HDPE and the biochar.
214
215 3.3 SEM
216 SEM images of HDPE, biochar and the fractured surfaces of biochar added composites are
217 shown in Figure 4. It is expected that the typical smooth structure of HDPE and the porous
218 structure of biochar were observed clearly by the SEM images. As can be observed in the SEM
219 image of B10, the biochar particles were dispersed in the HDPE matrix, but the dispersion was not
220 uniform. This can be attributed to an imbalance of the ratio between biochar and HDPE. For

5
221 further characterization, the dispersion gradually improved as the increase of biochar loadings,
222 whereas B70 showed poor dispersion and agglomeration due to the overloaded biochar. The
223 effects of biochar loadings on the microstructure were consistent with those of mechanical
224 properties. As stated earlier, the physical/mechanical interlocking structure caused by porous
225 biochar was also displayed clearly. It is evident that the penetration of the HDPE matrix in biochar
226 pores and cavitation caused by the HDPE matrix pullout shown in the images transferred the stress
227 effectively. Based on these, good dispersion and unique interface structure contributed to better
228 mechanical properties of biochar added composites (Zeng et al., 2018). The agglomeration and
229 cracks shown in the B70 image also gave an explanation for the high water absorption and poor
230 mechanical properties of B70.
231
232 3.4 DSC
233 The results obtained from the DSC analysis of HDPE and composite samples are illustrated
234 in Figure 5. In order to clearly reveal the effects of biochar loadings on the thermal properties of
235 composite samples, the DSC curves of different samples showed in Figure 5. Looking at the DSC
236 curves of HDPE, two main peaks were observed. The first peak around 119.78 °C and the second
237 peak around 130.80 °C are attributed to exothermic crystallization and endothermic melting of
238 HDPE respectively (Pandey et al., 2017). What is interesting is that the DSC curves of biochar
239 added composite samples showed significant differences. All the crystallization temperatures of
240 the composite samples appeared to be higher than that of HDPE. The higher the biochar loading,
241 the higher the crystallization temperature. Biochar acted as the crystal growth point as a nucleating
242 agent for crystallization of the HDPE with the early onset of exothermic crystallization occurred
243 (Das et al., 2016c), although the addition of biochar reduced the Xc of HDPE shown in Table 1 and
244 Figure 2. However, all the melting temperatures of the composites were slightly lower compared
245 to neat HDPE indicating a decreasing lamella thickness and smaller crystals (Li et al., 2016).
246 Moreover, the increase of energy required for melting in biochar added composites was observed
247 (Das et al., 2018). The increase of biochar loadings resulted in the increase of the required energy,
248 suggesting that the addition of biochar improved the thermal stability of HDPE.
249
250 3.5 TGA
251 The TG and DTG curves of HDPE and biochar added composites are demonstrated in Figure
252 6. Figure 6 (a) shows the mass loss curves and Figure 6 (b) shows the derivatives of the mass loss
253 curves. It can be observed from Figure 6 that the degradation of HDPE started at about 450 °C and
254 tended to be stable after 500 °C, the most intense degradation occurred at about 480 °C, and this
255 result belongs to the typical thermal decomposition process of HDPE (Lin et al., 2019).
256 Comparing with HDPE, all these three temperatures for all biochar added composites showed a
257 delay. And this delay was more significant in higher loadings of biochar added composites
258 especially for B70. On the other hand, the residue produced after the degradation of all samples
259 also showed a different trend. As shown in Figure. 6 (a), there was little residue produced from
260 HDPE after the temperature cycle indicating an almost complete degradation. It should be noted
261 that the biochar loading was positively correlated with the amount of residue produced, and the
262 comparison was as follows: B70 > B60> B50> B40> B30> B20> B10> HDPE. B70 retained the
263 maximum amount of residue (~60%). The proximate analysis from Figure S1 showed a high ash
264 content (~43.45%) in the biochar sample and a high SiO2 content (~89.5 wt%) in the ash. The

6
265 enrichment of stable SiO2 in biochar should be responsible for the amount of residue produced
266 (De Bhowmick et al., 2018). Moreover, the increase of the stable biochar was the main reason for
267 the increase of the amount of residue produced (Das et al., 2016a). In conclusion, biochar can
268 enhance the thermal property of HDPE in terms of the delay of decomposition temperatures.
269 Biochar shows great potential and benefits in the preparation of high thermal stability materials.
270
271 3.6 Viscoelastic behavior
272 Figure 7 presents the viscoelastic behavior including a variation of storage modulus and loss
273 factor with temperatures of HDPE and biochar added composites. According to Figure 7 (a),
274 increasing temperatures from -50 to 150 °C resulted in decreased storage modulus of all samples.
275 Since the increase of the temperature aggravated the thermal movement of HDPE molecules, the
276 deformation of composites resulted in poor dimensional stability (Pillai and Renneckar, 2016). In
277 addition, it is obvious that all the biochar added composites showed higher storage modulus than
278 that of HDPE suggesting that the stiffness of HDPE was improved with the addition of biochars.
279 As stated earlier, rigid biochar particles resulted in an increase of stiffness. The large difference
280 between the static mechanical properties and dynamic mechanical properties was that B70 showed
281 higher storage modulus among the samples. Subsequently, a variation of loss factor with
282 temperatures of all samples was shown in Figure 7 (b). It should be noted that all the biochar
283 added composites showed lower loss factor values than that of neat HDPE, which was attributed
284 to good interfacial interactions between biochar and HDPE. The decrease of the composites
285 caused by biochar in loss factor also indicated an improvement of HDPE elasticity.
286
287 3.7 Creep behavior and stress relaxation
288 Creep behavior and stress relaxation reflect the dimensional stability of a composite, and also
289 can be used to characterize the mechanical properties of a composite. As shown in Figure 8, the
290 creep behavior and stress relaxation of neat HDPE and biochar added composites were
291 characterized using creep compliance and stress relaxation modulus in this study. Both creep
292 behavior and stress relaxation of all samples showed three main stages suggesting the properties of
293 typical thermoplastic polymers. Interestingly, all the biochar added composites showed lower
294 creep compliance than that of HDPE, and the creep compliance continuously decreased with an
295 increase of the biochar amount in the HDPE composites. It can be stated that the addition of
296 biochar improved the creep resistance of HDPE. This can be attributed to the improvement of
297 dimensional stability, and the addition of rigid biochar improved the stiffness and increased the
298 ability to resist the deformation of the composites (Davis et al., 2019). Likewise, the stress
299 relaxation modulus continuously increased with an increase of the biochar amount in the HDPE
300 composites, and this suggested that the addition of biochar improved the anti-stress relaxation
301 ability of HDPE. Having better rigidity and a special porous structure should be responsible for
302 the enhancement of creep resistance and anti-stress relaxation ability. Hence, biochar exhibited
303 critical benefits in the mechanical properties of composites.
304
305 3.8 Flexural properties
306 Figure 9 compares the flexural properties including flexural strength and flexural modulus of
307 HDPE and biochar added composites. It can be observed that there was a significant difference in
308 HDPE with the addition of biochar in terms of flexural properties. Improvement of both flexural

7
309 strength and flexural modulus was achieved on the whole. The flexural strength and modulus of
310 HDPE were 14.14 MPa and 0.88 GPa, respectively whereas those of B10 were 16.37 MPa and
311 0.91 GPa, respectively. And this increment of flexural properties was reached due to the unique
312 characteristics of biochar. Having a high surface area (297.36 m2/g) and a special porous structure
313 should be responsible for the improvement of flexural properties. Das et al (Das et al., 2016a)
314 reported that at high temperatures, HPDE, which was in a fluid state, could fill the pores of the
315 biochar under the extrusion of the screw, and a physical/mechanical interlocking was resulted after
316 cooling. Since this physical/mechanical interlocking can transfer stress efficiently, it was the main
317 contribution for the better flexural properties. It can also be observed from Figure 9 that flexural
318 strength and modulus of samples B50 and B40 reached a maximum of 34.95 MPa and 1.79 GPa,
319 respectively. As stated earlier in Table S1, the high SiO2 content (~89.5 wt%) in the ash of biochar
320 was beneficial to improve the mechanical properties of the composites. Bulota et al. obtained a
321 similar result in PLA/algae composites (Bulota and Budtova, 2015). On the other hand, this
322 particular good increase can also be attributed to the good dispersion of biochar in HDPE as
323 shown in Figure S3. Many documents (Goud et al., 2019; Oliveira et al., 2018) confirmed that the
324 particle dispersion of fillers in the matrix was important for the flexural strength of a composite.
325 Also, the flexural properties of B60 and B70 confirmed this theory exactly. From B40 and B50 to
326 B60 and B70, the flexural properties of biochar added composites showed a significant reduction.
327 The most important reason for this decrease was the poor particle dispersion in the matrix. Less
328 HDPE matrix cannot effectively load more biochar particles, resulting in serious particle
329 agglomeration (Aliotta et al., 2019) and a compromise of flexural properties.
330
331 3.9 Tensile properties
332 Figure 10 compares the tensile properties including tensile strength and tensile modulus of
333 HDPE and biochar added composite samples. Determining the tensile properties of a composite
334 dictates its applications, and an overall improvement can be observed from Figure 10 with the
335 addition of biochar. The tensile strength of HDPE, B10, B20, B30, B40, B50, B60 and B70
336 samples were 23.54, 23.21, 24.88, 25.20, 29.05, 26.25, 24.90 and 19.02 MPa, respectively. When
337 the B40 biochar loading reached 40 wt%, the tensile strength increased by 23.41%, from 23.54
338 MPa for the neat HDPE to 29.05 MPa, and the maximum tensile strength among the samples was
339 obtained. In addition, the tensile modulus of HDPE, B10, B20, B30, B40, B50, B60 and B70
340 samples were 1.33, 1.36, 1.47, 1.69, 1.70, 1.87, 2.03 and 0.98 GPa, respectively. Also, B60
341 illustrated the maximum tensile modulus with a significant increase by 52.63%, from 1.33 GPa for
342 the neat HDPE to 2.03 GPa. Li et al. (Li et al., 2018) have summarized that better tensile strength
343 of a composite can be obtained by homogenous dispersion of filler particles in a polymer matrix,
344 which could give an explanation for better tensile strength of B40 and B50. Also, the stress
345 transfer mechanism caused by interfacial bonding dictates the tensile strength of a composite (Fu
346 et al., 2008). Thus, the betterment of the tensile strength in biochar added composites was strongly
347 dependent on the aforementioned special interface bonding. The increase of tensile modulus in
348 biochar added composites suggests another advantage in mechanics, and this improvement can be
349 attributed to the rigidity of biochars. The incorporation of rigid biochars into the HDPE matrix
350 reduced the deformability and mobility of HDPE macromolecules (Siebert and Wilker, 2019),
351 which improved the stiffness of biochar added composites. Moreover, biochar obtained at 600 °C
352 carbonization showed low hydrophilicity in this study, and this was confirmed by the contact

8
353 angle in Figure S2. Similar polarities resulting in good affinity could also be responsible for the
354 improvement of tensile properties of biochar added composites. Nevertheless, the tensile
355 properties of B70 decreased obviously. As can be known from Figure 1 that B70 was severely
356 deformed. Limited HDPE matrix cannot afford excessive biochar particles and the plasticity of the
357 composites was reduced. Excessive biochar caused the agglomeration, resulting in the fact that the
358 interfacial bonding strength of B70 was weakened, stress propagation was obstructed, and the
359 tensile properties of B70 were decreased (Nasser et al., 2019).
360
361 3.10 Water absorption
362 In addition to the above mechanical properties, water absorption is also an important factor
363 affecting the dimensional stability of a composite. The results of water absorption for HDPE and
364 biochar added composites in 28 days at room temperature are presented in Figure 11. According to
365 the figure, it is expected that the water absorption of all the samples increased with the immersion
366 time indicating the damage of dimensional stability by long-time immersion. Besides, HDPE
367 showed lower water absorption below 0.2%, this can be attributed to the microcracks produced in
368 the manufacturing process. Upon closer observation in Figure 11, it can be observed that the water
369 absorption increased for the biochar added composites as the biochar loading increased, and there
370 was a positive correlation between water absorption and biochar loading. This result exhibited the
371 worse water-resistance of biochar added composites than that of neat HDPE with the addition of
372 biochar. The water absorption of a composite is dependent on some factors such as fine pores,
373 lumens, and the gaps in the interface (Zabihzadeh, 2010). On the one hand, some interfacial voids
374 between biochar and HDPE were produced during the composite making process, and water was
375 absorbed by these interfacial voids (Guo et al., 2019). On the other hand, the porous structure of
376 biochar should also be responsible for the poor water resistance, although biochar showed low
377 hydrophilicity because of the decrease of hydroxyl groups under high temperatures. Additionally,
378 there was a significant difference in water absorption between B70 and other samples, and B70
379 showed the worst water resistance. This result can be attributed to the agglomeration of biochar in
380 HDPE and many biochar particles agglomerated together, which increased the contact area with
381 water and resulted in higher water absorption.
382
383 3.11 Flame retardancy
384 Since the flammability of a composite dictates application in outdoor, limited oxygen index
385 measurements were performed to characterize flame retardancy of HDPE and biochar added
386 composites in this study. Limited oxygen index refers to the volume ratio of oxygen in the mixture
387 of oxygen and nitrogen., The greater the limited oxygen index, the lower the flammability and the
388 better the flame retardant effect. As shown in Figure 12, HDPE shows a lower limited oxygen
389 index of 18.21% indicating its poor flame retardant, and this result was consistent with the report
390 of Jiang et al (Jiang et al., 2018). It should be noted that the limited oxygen index increased for the
391 biochar added composites as the biochar loading increased, which indicates that an increase in the
392 flame retardancy was obtained by adding biochar to HDPE. As stated earlier, the addition of
393 biochar improved the thermal properties of HDPE due to its high thermal stability. Thus, it can be
394 stated that the improvement of flame retardancy in HDPE was obtained by the addition of biochar,
395 and this can be attributed to that the biochar with high thermal stability hindered the transport of
396 heat between the heat source and HPDE (Das et al., 2016a). On the other hand, many documents

9
397 have reported that the preparation of biochar was accompanied by the production of a large
398 number of metal oxides and inorganic substances in the process of carbonization (Chen et al.,
399 2019; Wei et al., 2018). These nonflammable metal oxides and inorganic substances acting as
400 flame retardants can effectively slow down the combustion of HDPE and improve the flame
401 retardancy of it. Besides, the greatest limited oxygen index of 25.06% was obtained in B70
402 indicating the potential of biochar in flame retardant of composites.
403
404 4. Conclusions
405 In order to maximize the utilization of agricultural wastes, the effect of high biochar loadings
406 on biochar added composites was investigated and interesting results were obtained in this study.
407 The addition of biochar did not change the crystal planes of HDPE and no chemical reaction
408 occurred between HDPE and the biochar. Biochar caused the early onset of exothermic
409 crystallization of biochar added composites by DSC and improvement of thermal properties in
410 biochar added composites was achieved in terms of the amount of residue produced by TGA.
411 Moreover, the addition of biochar improved the mechanical properties of HDPE. Best flexural
412 strength of 34.95 MPa in B50, the flexural modulus of 1.79 GPa in B40, the tensile strength of
413 29.05 MPa in B40, and the tensile modulus of 2.03 GPa were obtained, which indicated that high
414 biochar loading filled composites with good mechanical properties were achieved. Besides, good
415 stiffness, elasticity, creep resistance and anti-stress relaxation ability were also obtained in high
416 biochar added composites. Although the increase of biochar loadings decreased the
417 water-resistance of the composites, the composites with highly filled biochars suggested better
418 flame retardant. The results presented here indicated that maximizing the utilization of agricultural
419 wastes to prepare green products with excellent properties can be achieved by highly loaded
420 biochar composites.
421

422 Acknowledgments
423 The authors appreciate that this work was financially supported by the Natural Science
424 Foundation of Shandong Province of China (No. ZR2019MEE036); the Agriculture and Food
425 Research Initiative Competitive Grant (No. 2016-67021-24533, and 2014-38502-22598) from the
426 National Institute of Food and Agriculture, United States Department of Agriculture; the National
427 Key Research and Development Program of China (No. 2018YFD1101001); the National Natural
428 Science Fund of China (No.51806129); the distinguished expert of Taishan scholars Shandong
429 province project and the higher education superior discipline team training program of Shandong
430 province.
431

432

433

434

435

10
436 References
437 Aliotta, L., Cinelli, P., Coltelli, M.B., Lazzeri, A., 2019. Rigid filler toughening in PLA-Calcium
438 Carbonate composites: Effect of particle surface treatment and matrix plasticization. European Polymer
439 Journal 113, 78-88.
440 Arnandha, Y., Satyarno, I., Awaludin, A., Irawati, I.S., Prasetya, Y., Prayitno, D.A., Winata, D.C., Satrio,
441 M.H., Amalia, A., 2017. Physical and mechanical properties of WPC board from sengon sawdust and
442 recycled HDPE plastic. Procedia engineering 171, 695-704.
443 Arrigo, R., Jagdale, P., Bartoli, M., Tagliaferro, A., Malucelli, G., 2019. Structure–property
444 relationships in polyethylene-based composites filled with biochar derived from waste coffee grounds.
445 Polymers 11, 1336.
446 Bajwa, D.S., Adhikari, S., Shojaeiarani, J., Bajwa, S.G., Pandey, P., Shanmugam, S.R., 2019.
447 Characterization of bio-carbon and ligno-cellulosic fiber reinforced bio-composites with compatibilizer.
448 Construction and Building Materials 204, 193-202.
449 Behazin, E., Misra, M., Mohanty, A.K., 2017. Sustainable biocarbon from pyrolyzed perennial grasses
450 and their effects on impact modified polypropylene biocomposites. Composites Part B: Engineering
451 118, 116-124.
452 Bulota, M., Budtova, T., 2015. PLA/algae composites: morphology and mechanical properties.
453 Composites Part A: Applied Science and Manufacturing 73, 109-115.
454 Chen, C., Yan, X., Xu, Y., Yoza, B.A., Wang, X., Kou, Y., Ye, H., Wang, Q., Li, Q.X., 2019. Activated
455 petroleum waste sludge biochar for efficient catalytic ozonation of refinery wastewater. Science of The
456 Total Environment 651, 2631-2640.
457 Chow, C.-F., Wong, W.-L., Chan, C.-W., Chan, C.-S., 2018. Converting inert plastic waste into
458 energetic materials: A study on the light-accelerated decomposition of plastic waste with the Fenton
459 reaction. Waste management 75, 174-180.
460 Das, O., Bhattacharyya, D., Hui, D., Lau, K.-T., 2016a. Mechanical and flammability characterisations
461 of biochar/polypropylene biocomposites. Composites Part B: Engineering 106, 120-128.
462 Das, O., Bhattacharyya, D., Sarmah, A.K., 2016b. Sustainable eco–composites obtained from waste
463 derived biochar: a consideration in performance properties, production costs, and environmental impact.
464 Journal of cleaner production 129, 159-168.
465 Das, O., Hedenqvist, M.S., Johansson, E., Olsson, R.T., Loho, T.A., Capezza, A.J., Raman, R.S.,
466 Holder, S., 2019. An all-gluten biocomposite: Comparisons with carbon black and pine char
467 composites. Composites Part A: Applied Science and Manufacturing 120, 42-48.
468 Das, O., Kim, N.K., Hedenqvist, M.S., Lin, R.J., Sarmah, A.K., Bhattacharyya, D., 2018. An attempt to
469 find a suitable biomass for biochar-based polypropylene biocomposites. Environmental management
470 62, 403-413.
471 Das, O., Kim, N.K., Kalamkarov, A.L., Sarmah, A.K., Bhattacharyya, D., 2017. Biochar to the rescue:
472 Balancing the fire performance and mechanical properties of polypropylene composites. Polymer
473 degradation and stability 144, 485-496.
474 Das, O., Sarmah, A.K., Bhattacharyya, D., 2015. A sustainable and resilient approach through biochar
475 addition in wood polymer composites. Science of the Total Environment 512, 326-336.
476 Das, O., Sarmah, A.K., Zujovic, Z., Bhattacharyya, D., 2016c. Characterisation of waste derived
477 biochar added biocomposites: chemical and thermal modifications. Science of the Total Environment
478 550, 133-142.

11
479 Davis, A.M., Hanzly, L.E., DeButts, B.L., Barone, J.R., 2019. Characterization of dimensional stability
480 in flax fiber reinforced polypropylene composites. Polymer Composites 40, 132-140.
481 De Bhowmick, G., Sarmah, A.K., Sen, R., 2018. Production and characterization of a value added
482 biochar mix using seaweed, rice husk and pine sawdust: A parametric study. Journal of Cleaner
483 Production 200, 641-656.
484 Demirbas, A., 2011. Waste management, waste resource facilities and waste conversion processes.
485 Energy Conversion and Management 52, 1280-1287.
486 Fa, W., Wang, J., Ge, S., Chao, C., 2019. Performance of photo-degradation and thermo-degradation of
487 polyethylene with photo-catalysts and thermo-oxidant additives. Polymer Bulletin, 1-16.
488 Ferreira, G.F., Pierozzi, M., Fingolo, A.C., da Silva, W.P., Strauss, M., 2019. Tuning Sugarcane
489 Bagasse Biochar into a Potential Carbon Black Substitute for Polyethylene Composites. Journal of
490 Polymers and the Environment 27, 1735-1745.
491 Fu, S.-Y., Feng, X.-Q., Lauke, B., Mai, Y.-W., 2008. Effects of particle size, particle/matrix interface
492 adhesion and particle loading on mechanical properties of particulate–polymer composites. Composites
493 Part B: Engineering 39, 933-961.
494 Giorcelli, M., Khan, A., Pugno, N.M., Rosso, C., Tagliaferro, A., 2019. Biochar as a cheap and
495 environmental friendly filler able to improve polymer mechanical properties. Biomass and bioenergy
496 120, 219-223.
497 Goud, V., Alagirusamy, R., Das, A., Kalyanasundaram, D., 2019. Influence of various forms of
498 polypropylene matrix (fiber, powder and film states) on the flexural strength of carbon-polypropylene
499 composites. Composites Part B: Engineering 166, 56-64.
500 Guo, G., Finkenstadt, V.L., Nimmagadda, Y., 2019. Mechanical properties and water absorption
501 behavior of injection-molded wood fiber/carbon fiber high-density polyethylene hybrid composites.
502 Advanced Composites and Hybrid Materials, 1-11.
503 Hahn, A., Gerdts, G., Völker, C., Niebühr, V., 2019. Using FTIRS as pre-screening method for
504 detection of microplastic in bulk sediment samples. Science of the total environment 689, 341-346.
505 Ho, M.-p., Lau, K.-t., Wang, H., Hui, D., 2015. Improvement on the properties of polylactic acid (PLA)
506 using bamboo charcoal particles. Composites Part B: Engineering 81, 14-25.
507 Jiang, D., Pan, M., Cai, X., Zhao, Y., 2018. Flame retardancy of rice straw-polyethylene composites
508 affected by in situ polymerization of ammonium polyphosphate/silica. Composites Part A: Applied
509 Science and Manufacturing 109, 1-9.
510 Li, S., Huang, A., Chen, Y.-J., Li, D., Turng, L.-S., 2018. Highly filled biochar/ultra-high molecular
511 weight polyethylene/linear low density polyethylene composites for high-performance electromagnetic
512 interference shielding. Composites Part B: Engineering 153, 277-284.
513 Li, S., Li, X., Chen, C., Wang, H., Deng, Q., Gong, M., Li, D., 2016. Development of electrically
514 conductive nano bamboo charcoal/ultra-high molecular weight polyethylene composites with a
515 segregated network. Composites Science and Technology 132, 31-37.
516 Lim, S.L., Lee, L.H., Wu, T.Y., 2016. Sustainability of using composting and vermicomposting
517 technologies for organic solid waste biotransformation: recent overview, greenhouse gases emissions
518 and economic analysis. Journal of Cleaner Production 111, 262-278.
519 Lin, X., Kong, L., Cai, H., Zhang, Q., Bi, D., Yi, W., 2019. Effects of alkali and alkaline earth metals
520 on the co-pyrolysis of cellulose and high density polyethylene using TGA and Py-GC/MS. Fuel
521 processing technology 191, 71-78.
522 Mashouf Roudsari, G., Mohanty, A.K., Misra, M., 2017. A statistical approach to develop

12
523 biocomposites from epoxy resin, poly (furfuryl alcohol), poly (propylene carbonate), and biochar.
524 Journal of Applied Polymer Science 134, 45307.
525 Mohanty, A.K., Vivekanandhan, S., Pin, J.-M., Misra, M., 2018. Composites from renewable and
526 sustainable resources: Challenges and innovations. Science 362, 536-542.
527 Nan, N., DeVallance, D.B., Xie, X., Wang, J., 2016. The effect of bio-carbon addition on the electrical,
528 mechanical, and thermal properties of polyvinyl alcohol/biochar composites. Journal of Composite
529 Materials 50, 1161-1168.
530 Nasser, J., Lin, J., Steinke, K., Sodano, H.A., 2019. Enhanced interfacial strength of aramid fiber
531 reinforced composites through adsorbed aramid nanofiber coatings. Composites Science and
532 Technology 174, 125-133.
533 Oliveira, L.Á., Santos, J.C., Panzera, T.H., Freire, R.T., Vieira, L.M., Scarpa, F., 2018. Evaluation of
534 hybrid-short-coir-fibre-reinforced composites via full factorial design. Composite Structures 202,
535 313-323.
536 Pandey, P., Bajwa, S.G., Bajwa, D.S., Englund, K., 2017. Performance of UV weathered HDPE
537 composites containing hull fiber from DDGS and corn grain. Industrial crops and products 107,
538 409-419.
539 Pillai, K.V., Renneckar, S., 2016. Dynamic mechanical analysis of layer-by-layer cellulose
540 nanocomposites. Industrial Crops and Products 93, 267-275.
541 Poulose, A.M., Elnour, A.Y., Anis, A., Shaikh, H., Al-Zahrani, S., George, J., Al-Wabel, M.I., Usman,
542 A.R., Ok, Y.S., Tsang, D.C., 2018. Date palm biochar-polymer composites: An investigation of
543 electrical, mechanical, thermal and rheological characteristics. Science of the total environment 619,
544 311-318.
545 Santhiago, M., Garcia, P.S., Strauss, M., 2018. Bio-based nanostructured carbons toward sustainable
546 technologies. Current Opinion in Green and Sustainable Chemistry 12, 22-26.
547 Siebert, H.M., Wilker, J.J., 2019. Deriving Commercial Level Adhesive Performance from a Bio-Based
548 Mussel Mimetic Polymer. ACS Sustainable Chemistry & Engineering 7, 13315-13323.
549 Song, B., Chen, M., Zhao, L., Qiu, H., Cao, X., 2019. Physicochemical property and colloidal stability
550 of micron-and nano-particle biochar derived from a variety of feedstock sources. Science of The Total
551 Environment 661, 685-695.
552 Turku, I., Keskisaari, A., Kärki, T., Puurtinen, A., Marttila, P., 2017. Characterization of wood plastic
553 composites manufactured from recycled plastic blends. Composite Structures 161, 469-476.
554 Wang, X., Peng, S., Chen, H., Yu, X., Zhao, X., 2019. Mechanical properties, rheological behaviors,
555 and phase morphologies of high-toughness PLA/PBAT blends by in-situ reactive compatibilization.
556 Composites Part B: Engineering, 107028.
557 Wei, D., Li, B., Huang, H., Luo, L., Zhang, J., Yang, Y., Guo, J., Tang, L., Zeng, G., Zhou, Y., 2018.
558 Biochar-based functional materials in the purification of agricultural wastewater: fabrication,
559 application and future research needs. Chemosphere 197, 165-180.
560 Zabihzadeh, S.M., 2010. Water uptake and flexural properties of natural filler/HDPE composites.
561 BioResources 5, 316-232.
562 Zeng, S., Shen, M., Yang, L., Xue, Y., Lu, F., Chen, S., 2018. Self-assembled montmorillonite–carbon
563 nanotube for epoxy composites with superior mechanical and thermal properties. Composites Science
564 and Technology 162, 131-139.
565 Zhang, Q., Khan, M.U., Lin, X., Cai, H., Lei, H., 2019a. Temperature varied biochar as a reinforcing
566 filler for high-density polyethylene composites. Composites Part B: Engineering 175, 107151.

13
567 Zhang, Q., Li, Y., Cai, H., Lin, X., Yi, W., Zhang, J., 2019b. Properties comparison of high density
568 polyethylene composites filled with three kinds of shell fibers. Results in Physics 12, 1542-1546.
569 Zhang, Q., Yi, W., Li, Z., Wang, L., Cai, H., 2018. Mechanical properties of rice husk biochar
570 reinforced high density polyethylene composites. Polymers 10, 286.
571

14
Figures

Fig. 1. Composites samples under different biochar loading


100

20

HDPE
B10
Intensity

B20
B30
B40
B50
B60
B70

10 20 30 40 50 60
2θ/(°)
Fig.
2. XRD curves of HDPE and composites samples
1080 1470 2850 2920 3400
HDPE

B10

B20
Abosorbance

B30

B40

B50
B60

B70

1000 1500 2000 2500 3000 3500


-1
Wavenumber(cm )
Fig.
3. FTIR spectra of HDPE and composites samples
Fig. 4. SEM images of HDPE, biochar and composites samples
119.78 °C
B70 130.80 °C

B60
B50
Heat flow(W/g)

B40

B30

B20

B10

HDPE

40 60 80 100 120 140 160

Temperature(°C)
Fig.
5. DSC curves of HDPE and composites samples
(a) 100

80

HDPE
Mass loss(%)

60
B10
B20
40 B30
B40
B50
20
B60
B70
0

100 200 300 400 500 600

Temperature(°C)

(b) 0

-10
Mass loss rate/(% min )
-1

HDPE
B10
-20
B20
B30
B40
-30
B50
B60
-40 B70

-50
100 200 300 400 500 600

Temperature(°C)

Fig. 6. TGA of HDPE and composites samples: (a) TG (b) DTG


3000

(a)
HDPE
2500
B10
B20
Storage Modulus(MPa)

2000 B30
B40
B50
1500
B60
B70
1000

500

-50 -25 0 25 50 75 100 125 150

Temperature(°C)

0.40

(b)
0.35 HDPE
B10
0.30 B20
B30
0.25 B40
Loss Factor

B50
0.20 B60
B70
0.15

0.10

0.05

0.00

-50 -25 0 25 50 75 100 125 150

Temperature(°C)

Fig. 7. Viscoelastic behavior of HDPE and composites samples: (a) storage modulus (b) loss factor
4000
(a)
3500

HDPE
Creep Compliance(μm /N)

3000
B10
2

B20
2500
B30
B40
2000 B50
B60
1500 B70

1000

500

5 10 15 20 25 30 35 40 45

Time(min)

1600
(b)
1400

HDPE
Relaxation Modulus(MPa)

1200
B10
1000 B20
B30
800
B40
B50
B60
600
B70

400

200

5 10 15 20 25 30 35 40 45

Time(min)

Fig. 8. Variation of (a) creep behavior and (b) stress relaxation in 30min under for HDPE and composites
samples
40
(a) 34.95
35

30 28.15 29.19
Flexural Strength(MPa)

25

19.33 18.09
20 18.3
16.37
14.16
15

10

0
HDPE B10 B20 B30 B40 B50 B60 B70

2.2
(b)
2.0
1.79 1.76
1.8

1.6
Flexural Modulus(GPa)

1.4
1.29 1.32 1.3
1.11
1.2

1.0
0.88 0.91
0.8

0.6

0.4

0.2

0.0
HDPE B10 B20 B30 B40 B50 B60 B70

Fig. 9. Flexural properties of HDPE and composites samples: (a) flexural strength (b) flexural modulus
(a) 29.05
30
26.25 24.9
23.54 23.21 24.88 25.2
25
Tensile Strength(MPa)

19.02
20

15

10

0
HDPE B10 B20 B30 B40 B50 B60 B70

(b)
2.03
1.87
2.0 1.7
1.69
1.33
Tensile Modulus(GPa)

1.47
1.36
1.5
0.98

1.0

0.5

0.0
HDPE B10 B20 B30 B40 B50 B60 B70

Fig. 10. Tensile properties of HDPE and composites samples: (a) tensile strength (b) tensile modulus
2.0

HDPE
1.5 B10
B20
Water Absorption(%)

B30
1.0
B40
B50
B60
B70
0.5

0.0

0 7 14 21 28
Time(d)

Fig. 11. Variation of water absorption in 28 days for HDPE and composites samples
25.06
25 23.91
21.69 22.17 22.44
Limited Oxygen Index(%)

19.93 20.17
20 18.21

15

10

0
HDPE B10 B20 B30 B40 B50 B60 B70

Fig. 12. Limited oxygen index of HDPE and composites samples


Table 1 Crystallization and melting behavior of HDPE and composites samples
Samples Tc (°C) Tm (°C) Xc (%)
HDPE 119.78 130.80 74.36
B10 119.83 130.96 60.15
B20 119.88 130.65 58.41
B30 120.04 130.54 59.12
B40 120.09 130.51 52.74
B50 120.22 130.48 50.61
B60 120.24 130.61 51.08
B70 120.53 130.24 47.44

You might also like