Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Industrial Crops & Products 146 (2020) 112185

Contents lists available at ScienceDirect

Industrial Crops & Products


journal homepage: www.elsevier.com/locate/indcrop

Biochar filled high-density polyethylene composites with excellent T


properties: Towards maximizing the utilization of agricultural wastes
Qingfa Zhanga,b, Donghong Zhanga, Hang Xua, Wenyu Lua, Xiajin Rena, Hongzhen Caia,*,
Hanwu Leib,*, Erguang Huob, Yunfeng Zhaob, Moriko Qianb, Xiaona Lina,b, Elmar M. Villotab,
Wendy Mateob
a
School of Agricultural Engineering and Food Science, Shandong Research Center of Engineering & Technology for Clean Energy, Shandong University of Technology, Zibo,
255000, China
b
Department of Biological Systems Engineering, Washington State University, Richland, WA 99354, USA

A R T I C LE I N FO A B S T R A C T

Keywords: Biochar derived from agricultural wastes was used to reinforce high-density polyethylene (HDPE) to obtain
Agricultural wastes composites abbreviated as B10, B20, B30, B40, B50, B60, and B70. Investigating the mechanical, thermal, water
Maximization absorption and flame retardant properties of the composites is one of the objectives while maximizing the
Biochar utilization of agricultural wastes is the ultimate goal of this work. It was found that flexural properties, tensile
Composites
properties, storage modulus, elasticity, creep resistance and anti-stress relaxation ability of HDPE were improved
Properties evaluation
by the inclusion of biochar, and excellent mechanical properties were obtained in 50 % even 60 % biochar added
composites because of good dispersion and unique structure shown in the interface. The composites achieved
good flexural strength of 34.95 MPa in B50, flexural modulus of 1.79 GPa in B40, tensile strength of 29.05 MPa
in B40, and tensile modulus of 2.03 GPa. Additionally, the thermal and flame retardant properties (limited
oxygen index of 25.06 % in B70) increased for the biochar added composites as the biochar loading increased
due to the high thermal stability of biochar, although biochar had a negative effect on water-resistance of the
composites. The results revealed that the ultimate goal was achieved in terms of producing composites with
excellent mechanical, thermal, water absorption and flame retardant properties while maximizing the utilization
of agricultural wastes as a rational balance.

1. Introduction wastes are renewable comparing with fossil fuels. In a nutshell, in-
cineration and discarding of agricultural wastes are the waste of re-
The demand for fossil fuels increases dramatically along with sources. On the other hand, polyolefine plastics like polyethylene (PE),
worldwide population growth, economic progress, and technological polypropylene (PP), polyvinyl chloride (PVC), etc. have become an ir-
advancement. One great challenge is the pollution caused by the fossil replaceable raw material for industrial production and daily life due to
fuel with its depletion becoming another problem that cannot be ig- their excellent characteristics with technological advancement (Chow
nored (Das et al., 2015). On the other hand, the urgent demand for food et al., 2018). Be that as it may, polyolefine plastics show non-degrad-
due to population growth has contributed to the agglomeration of a ability which is responsible for serious environmental pollution. Like-
large number of agricultural wastes. In traditional agricultural pro- wise, the main disposals of waste plastics, such as incineration and
duction, incineration and discarding are the main ways in disposing of landfilling, have created more serious problems of polluting the air and
agricultural wastes (Demirbas, 2011). Nevertheless, these two ways water seriously (Fa et al., 2019). The recycling of waste plastics has
pollute the air and water and aggravate the greenhouse effect showing a become another great problem affecting environmental safety. Given all
huge threat to agricultural product security and environmental safety that, the effective and practical disposal of agricultural and plastic
(Lim et al., 2016). Hence, the effective disposal of agricultural wastes is wastes is a real challenge especially with the current pressure of efforts
an urgent need. In fact, agricultural wastes have great potential for on environmental protection.
resource utilization due to their rich organic matters including cellu- In the view of efficiently utilizing plastics and agricultural wastes,
lose, hemicellulose, lignin, etc. Far more important is that agricultural biocomposites made great progress worldwide, especially in North


Corresponding authors.
E-mail addresses: chzh666666@126.com (H. Cai), hlei@wsu.edu (H. Lei).

https://doi.org/10.1016/j.indcrop.2020.112185
Received 30 October 2019; Received in revised form 28 December 2019; Accepted 29 January 2020
0926-6690/ © 2020 Elsevier B.V. All rights reserved.
Q. Zhang, et al. Industrial Crops & Products 146 (2020) 112185

America over the past decades. Normally, some biocomposites are biochar loadings. A balance of biocomposites properties and economic
prepared with plastic wastes like PE, PP, PVC, etc. and agricultural efficiency is expected to achieve in terms of maximizing the use of
wastes like corn stalk, rice husk, peanut shell, etc. by extrusion, hot agricultural wastes while minimizing the production cost.
pressing and injection molding (Arrigo et al., 2019; Ferreira et al.,
2019; Zhang et al., 2019b). The biocomposites have been used in 2. Materials and methods
construction, transportation, industry, and decoration due to their
corrosion resistance, easy processing, and acid and alkali resistance. 2.1. The constituents
Importantly, the biocomposites can be recycled and reused discharging
no pollution in the production process (Wang et al., 2019). Never- Rice husk was used as feedstock to produce biochar at 600 °C using a
theless, because of the incompatibility between agricultural waste fi- miniature box furnace (KSL-1100X-S). To get a higher yield, the pyr-
bers and plastics, the interfacial adhesion of these biocomposites is olysis temperature was increased from 25 °C (ambient temperature) to
typically weak causing the poor mechanical properties subsequently 600 °C at a low heating rate (5 °C/min) under 20 mL/min nitrogen at-
(Turku et al., 2017). To overcome this issue, interfacial compatibilizers mosphere and held for 2 h to obtain the biochar sample. The biochar
such as maleic anhydride are added to the biocomposites and the me- sample was sieved over 150 μm screen to keep biochar particle size
chanical properties are efficiently improved. Bajwa et al. studied the below150 μm. The main parameters of the biochar sample and ashes
effect of maleic anhydride grafting polyethylene (MAPE) on the prop- were shown in Table S1. HDPE (9001) pellets with a melt index of
erties of high-density polyethylene (HDPE)/wood fiber composites, and 0.05 g/10 min and a density of 0.950 g/cm3 were purchased from USI
better mechanical properties were obtained with the addition of MAPE Corporation. The biochar and HDPE were dried at 105 °C and 50 °C for
(Bajwa et al., 2019). Nevertheless, the addition of costly interfacial 24 h prior to compounding with each other.
compatibilizers increased the production cost of biocomposites in-
dicating a new problem. The balance of properties and the economic 2.2. Composites processing
efficiency of biocomposites has become a new challenge. Therefore, lots
of studies have been done to solve this problem. Among them, the Dried biochar particles and HDPE pellets were put into a high-speed
utilization of biochar for biocomposites attracted extensive attention mixer for 10 min to get the uniform blends. Subsequently, a micro twin-
due to its porous structure and large surface area in recent years (Das screw extruder (WLG10 G) was used to melt and mix the composites
et al., 2015; Nan et al., 2016). blends, and then a micro-injection molding machine (WZS10D) was
All the agricultural wastes could be converted to biochar by a used to get cuboid shaped and dog bone shaped composite samples. The
thermochemical process known as carbonization or pyrolysis with extruding zone and blending zone of the extruder were run at 30 rpm
abundance feedstock sources, which lower the production cost of bio- and 180 °C, the four temperature zones of injection moulding were all
char. Moreover, biochar reinforced biocomposites achieve the balance set to 180 °C and the mold was kept at 50 °C. To maximize the utili-
between properties and cost (Behazin et al., 2017; Das et al., 2019). On zation of agricultural wastes, increasing the loading of biochar as much
the one hand, a physical/mechanical interlocking structure (Das et al., as possible is necessary for this study. It was found that 70 wt% loading
2017) formed by porous biochar in plastics effectively transfers stress of biochar was the maximum load that can be afforded in the compo-
and improves the mechanical properties of biocomposites. Bajwa et al. sites in view of molding and deformation as shown in Fig. 1. Based on
and Das et al. found that the coupling agent was not effective in biochar this, the biochar/HDPE mixtures contained biochar ranging from 10 to
added composites indicating unnecessary of using interfacial compati- 70 wt% with an interval of 10 wt% each, and the composite samples
bilizers which reduced the cost (Bajwa et al., 2019; Das et al., 2016b). under different biochar loadings were named as B10, B20, B30, B40,
Subsequently, lots of efforts have been done in biochar biocomposites in B50, B60, and B70 respectively.
terms of the unique advantages of biochar. The study of Giorcelli et al.
showed that biochar was a cheap and environmental friendly filler to 2.3. Composites characterization
improve polymer mechanical properties (Giorcelli et al., 2019). Ho
et al. used bamboo biochar to reinforce polylactic acid (PLA), and an Chemical characterization of HDPE and different biocomposites was
excellent improvement in the properties of PLA was obtained (Ho et al., conducted using Fourier transform infrared spectroscopy (FTIR) and X-
2015). Das et al. thought that the utilization of biochar to prepare Ray diffraction (XRD). FTIR of the samples was performed in a Fourier
biocomposites was a sustainable and resilient approach to manage transform infrared spectroscopy (Nicolet 5700, Thermo Fisher Nicolet,
agricultural wastes (Das et al., 2015). The development of biochar USA) with a scanning range from 4000 cm−1 to 400 cm−1. XRD of the
biocomposites attracted more and more scholars to participate in the samples was performed in a polycrystalline X-ray diffractometer
study, and the number of available documents has been more than 50 in
the last five years. The biochar sources (Poulose et al., 2018; Santhiago
et al., 2018; Zhang et al., 2018) are not limited to agricultural wastes
involving forestry wastes, grasses, manure, even bones, etc. Recently,
Mohanty et al. published a paper in Science demonstrating that biochar
was a new sustainable and functional filler material in biocomposites
(Mohanty et al., 2018). Besides, the effects of biochar loading, biochar
feedstock sources and types of polymers on biocomposites are also
studied. Though researchers have a universal agreement that biochar
loading has the greatest influence on biochar biocomposites, the bio-
char loading of most investigations is low. The maximum biochar
loadings were 35 % (Das et al., 2016a), 10 % (Ho et al., 2015), 20 %
(Mashouf Roudsari et al., 2017) shown by available documents re-
spectively. In view of production cost of polymers, it is necessary to
increase the biochar loading in biocomposites. High biochar loadings
help maximize the utilization of agricultural wastes on the premise of
ensuring environmental safety and reducing costs. To date, the effect of
high biochar loadings on biochar biocomposites is rarely investigated.
Hence, the focus of this work is to optimize biocomposites by regulating Fig. 1. Composites samples under different biochar loadings.

2
Q. Zhang, et al. Industrial Crops & Products 146 (2020) 112185

(Bruker AXS D8 Advance, Germany) using CuKα radiation with 2θ


varying between 5° and 60° at 5 °/min. Microstructure characterization
of fractured surfaces of different biocomposites was conducted in a
scanning electron microscope (SEM) (FEI Sirion 200, USA) with a
scanning voltage of 3.0 kV.
Thermal characterization of HDPE and different biocomposites was
conducted using Differential scanning calorimetry (DSC) and thermo-
gravimetric analysis (TGA). DSC of the samples was performed in a
differential scanning calorimeter analyzer (DSC-Q100, TA Instrument,
USA), and all the samples were heated to 180 °C and then cooled to
100 °C and then heated to 180 °C, all at 5 °C/min. TGA of the samples
was performed in a synchronous thermal analyzer (STA 449,
NETZSCH), and all the samples were heated from 30 to 600 °C under
the 20 mL/min nitrogen atmosphere with a heating rate of 10 °C/min.
Mechanical characterization of HDPE and different biocomposites
was conducted by dynamic mechanical analysis (DMA) and static me-
chanical analysis (SMA). DMA including viscoelastic behavior, creep Fig. 2. XRD spectra of HDPE and composites samples.
compliance and relaxation modulus was performed using a DMA Q 800
(TA instrument, USA). The samples were heated from −50 to 150 °C to
obtain the viscoelastic behavior in a single cantilever mode with the Table 1
Crystallization and melting behavior of HDPE and composites samples.
heating rate of 5 °C/min, and the 30 min creep compliance and re-
laxation modulus of the samples were obtained using the Creep TTS and Samples Tc (°C) Tm (°C) Xc (%)
Stress Relaxation TTS models of DMA under 30 °C. SMA including
HDPE 119.78 130.80 74.36
flexural and tensile properties was performed using an electronic uni- B10 119.83 130.96 60.15
versal testing machine (WDW1020, Changchun Kexin Co., Ltd., China). B20 119.88 130.65 58.41
Five specimens of each composite sample were tested. The average B30 120.04 130.54 59.12
values were reported with their standard errors, and the results were B40 120.09 130.51 52.74
B50 120.22 130.48 50.61
analyzed with SPSS using a one-way analysis of variance (ANOVA). The
B60 120.24 130.61 51.08
statistical analysis showed that the composites with varied loadings of B70 120.53 130.24 47.44
biochar had a significant difference in flexural and tensile properties
(p < 0.05).
HDPE and different biocomposites were completely immersed in biochar did not change the crystal planes of HDPE. A big intensity drop
water at ambient temperature for durations of up to 4 weeks, and the was observed in B10, which was due to that biochar addition even at
samples were removed and wiped to dry to measure the weights using a low loading (10 %) reduced the crystallinity degree (Xc) of HDPE as
precision weigh balance. Statistical analysis showed that the composites shown in Table 1. Additionally, the intensity of these two peaks in
with varied biochar loadings had a significant difference in water ab- composites was reduced as the loading of biochar increased. Since the
sorption (p < 0.05). The water uptake of samples was calculated by: biochar is amorphous in nature (Das et al., 2015), the increment of
amorphous biochar and reduction of Xc with the increase of biochar
Water Absorption (%) = (W1 − W0) / W0 × 100 %
loadings should be responsible for the reduction of major peaks in
where W1 and W0 are the weights of before and after soaking respec- composites (Zhang et al., 2019a).
tively.
The flame retardancy of HDPE and different biocomposites was
3.2. FTIR
characterized by a limited oxygen index using an oxygen index tester
(ZR-01, China), and the flow rate of oxygen and oxygen-nitrogen mix-
Fig. 3 illustrates the FTIR spectra of HDPE and composite samples.
ture were 3 L/min and 20 L/min, respectively. Statistical analysis
As can be observed from Fig. 3 that three typical peaks at 2920, 2850
showed that the composites with varied biochar loadings had a sig-
and 1470 cm−1 were shown in the FTIR spectra of HDPE and these
nificant difference in the limited oxygen index (p < 0.05). The limited
oxygen index of samples was calculated by:

Limited Oxygen Index (%) = [O2] / ([O2] + [N2]) × 100 %

where [O2] and [N2] are the flow rate of oxygen and nitrogen, re-
spectively

3. Results and discussion

3.1. XRD

Fig. 2 provides the results obtained from the XRD analysis of HDPE
and composite samples. From the figure, we can see two typical char-
acteristic peaks at 2θ of 21.6° and 24° which corresponded to (110) and
(200) crystal planes. Many documents confirmed that these two typical
crystal planes belonged to neat PE (Arnandha et al., 2017). It should be
noted that the position of the two typical characteristic peaks was not
changed with the addition of loading varied biochars indicating that the
two major peaks were both contributed by HDPE and the addition of Fig. 3. FTIR spectra of HDPE and composites samples.

3
Q. Zhang, et al. Industrial Crops & Products 146 (2020) 112185

three peaks can be assigned to the C–H stretching vibration of aliphatic


structures (Hahn et al., 2019). Highly biochar added composites,
especially B10, showed weaker peaks at 2920, 2850 and 1470 cm−1,
which was attributed to the lack of HDPE in composites. Interestingly,
peaks involving asymmetrical OeH stretching vibration and CeO
stretching vibration around 3400 cm−1 and 1080 cm−1 were weakly
detected in composites samples (Song et al., 2019). In addition, both
two peaks in the composites were contributed by the biochar. The
asymmetrical OeH stretching vibration was from phenolic hydroxyl
and alcohol hydroxyl in rice husk, and CeO stretching vibration was
caused by carbohydrates. The weakening of these two functional groups
indicates that the carbonization at 600 °C reduced the polarity of rice
husk and resulted in a low hydrophilicity. The similar results were also
obtained in Fig. S2. The contact angle analysis showed that biochar
added composites presented lower hydrophilicity compared to rice
husk/HDPE composites, which indicates that biochar added composites
had better mechanical properties than rice husk/HDPE composites. The
document demonstrated that the coupling agent showed little effect in Fig. 5. DSC spectra of HDPE and composites samples.
biochar/HDPE composites, which confirmed the analysis of this study
(Bajwa et al., 2019). Moreover, changes in the peak intensities, shifts due to the overloaded biochar. The effects of biochar loadings on the
and appearances or disappearances were not observed among the microstructure were consistent with those of mechanical properties. As
samples, which was evident that no chemical reaction occurred be- stated earlier, the physical/mechanical interlocking structure caused by
tween HDPE and the biochar. porous biochar was also displayed clearly. It is evident that the pene-
tration of the HDPE matrix in biochar pores and cavitation caused by
the HDPE matrix pullout shown in the images transferred the stress
3.3. SEM effectively. Based on these, good dispersion and unique interface
structure contributed to better mechanical properties of biochar added
SEM images of HDPE, biochar and the fractured surfaces of biochar composites (Zeng et al., 2018). The agglomeration and cracks shown in
added composites are shown in Fig. 4. It is expected that the typical the B70 image also gave an explanation for the high water absorption
smooth structure of HDPE and the porous structure of biochar were and poor mechanical properties of B70.
observed clearly by the SEM images. As can be observed in the SEM
image of B10, the biochar particles were dispersed in the HDPE matrix,
but the dispersion was not uniform. This can be attributed to an im- 3.4. DSC
balance of the ratio between biochar and HDPE. For further char-
acterization, the dispersion gradually improved as the increase of bio- The results obtained from the DSC analysis of HDPE and composite
char loadings, whereas B70 showed poor dispersion and agglomeration samples are illustrated in Fig. 5. In order to clearly reveal the effects of

Fig. 4. SEM images of HDPE, biochar and composites samples.

4
Q. Zhang, et al. Industrial Crops & Products 146 (2020) 112185

biochar loadings on the thermal properties of composite samples, the decomposition process of HDPE (Lin et al., 2019). Comparing with
DSC curves of different samples showed in Fig. 5. Looking at the DSC HDPE, all these three temperatures for all biochar added composites
curves of HDPE, two main peaks were observed. The first peak around showed a delay. And this delay was more significant in higher loadings
119.78 °C and the second peak around 130.80 °C are attributed to of biochar added composites especially for B70. On the other hand, the
exothermic crystallization and endothermic melting of HDPE respec- residue produced after the degradation of all samples also showed a
tively (Pandey et al., 2017). What is interesting is that the DSC curves of different trend. As shown in Fig. 6(a), there was little residue produced
biochar added composite samples showed significant differences. All from HDPE after the temperature cycle indicating an almost complete
the crystallization temperatures of the composite samples appeared to degradation. It should be noted that the biochar loading was positively
be higher than that of HDPE. The higher the biochar loading, the higher correlated with the amount of residue produced, and the comparison
the crystallization temperature. Biochar acted as the crystal growth was as follows: B70 > B60 > B50 > B40 > B30 > B20 > B10 >
point as a nucleating agent for crystallization of the HDPE with the HDPE. B70 retained the maximum amount of residue (∼60 %). The
early onset of exothermic crystallization occurred (Das et al., 2016c), proximate analysis from Fig. S1 showed a high ash content (∼43.45 %)
although the addition of biochar reduced the Xc of HDPE shown in in the biochar sample and a high SiO2 content (∼89.5 wt%) in the ash.
Table 1 and Fig. 2. However, all the melting temperatures of the The enrichment of stable SiO2 in biochar should be responsible for the
composites were slightly lower compared to neat HDPE indicating a amount of residue produced (De Bhowmick et al., 2018). Moreover, the
decreasing lamella thickness and smaller crystals (Li et al., 2016). increase of the stable biochar was the main reason for the increase of
Moreover, the increase of energy required for melting in biochar added the amount of residue produced (Das et al., 2016a). In conclusion,
composites was observed (Das et al., 2018). The increase of biochar biochar can enhance the thermal property of HDPE in terms of the delay
loadings resulted in the increase of the required energy, suggesting that of decomposition temperatures. Biochar shows great potential and
the addition of biochar improved the thermal stability of HDPE. benefits in the preparation of high thermal stability materials.

3.5. TGA 3.6. Viscoelastic behavior

The TG and DTG curves of HDPE and biochar added composites are Fig. 7 presents the viscoelastic behavior including a variation of
demonstrated in Fig. 6. Fig. 6(a) shows the mass loss curves and storage modulus and loss factor with temperatures of HDPE and biochar
Fig. 6(b) shows the derivatives of the mass loss curves. It can be ob- added composites. According to Fig. 7(a), increasing temperatures from
served from Fig. 6 that the degradation of HDPE started at about 450 °C −50 to 150 °C resulted in decreased storage modulus of all samples.
and tended to be stable after 500 °C, the most intense degradation oc- Since the increase of the temperature aggravated the thermal move-
curred at about 480 °C, and this result belongs to the typical thermal ment of HDPE molecules, the deformation of composites resulted in

Fig. 7. Viscoelastic behavior of HDPE and composites samples: (a) storage


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

5
Q. Zhang, et al. Industrial Crops & Products 146 (2020) 112185

Fig. 8. Variation of (a) creep behavior and (b) stress relaxation in 30 min for
HDPE and composites samples. Fig. 9. Flexural properties of HDPE and composites samples: (a) flexural
strength (b) flexural modulus.

poor dimensional stability (Pillai and Renneckar, 2016). In addition, it


is obvious that all the biochar added composites showed higher storage biochar improved the stiffness and increased the ability to resist the
modulus than that of HDPE suggesting that the stiffness of HDPE was deformation of the composites (Davis et al., 2019). Likewise, the stress
improved with the addition of biochars. As stated earlier, rigid biochar relaxation modulus continuously increased with an increase of the
particles resulted in an increase of stiffness. The large difference be- biochar amount in the HDPE composites, and this suggested that the
tween the static mechanical properties and dynamic mechanical prop- addition of biochar improved the anti-stress relaxation ability of HDPE.
erties was that B70 showed higher storage modulus among the samples. Having better rigidity and a special porous structure should be re-
Subsequently, a variation of loss factor with temperatures of all samples sponsible for the enhancement of creep resistance and anti-stress re-
was shown in Fig. 7(b). It should be noted that all the biochar added laxation ability. Hence, biochar exhibited critical benefits in the me-
composites showed lower loss factor values than that of neat HDPE, chanical properties of composites.
which was attributed to good interfacial interactions between biochar
and HDPE. The decrease of the composites caused by biochar in loss 3.8. Flexural properties
factor also indicated an improvement of HDPE elasticity.
Fig. 9 compares the flexural properties including flexural strength
3.7. Creep behavior and stress relaxation and flexural modulus of HDPE and biochar added composites. It can be
observed that there was a significant difference in HDPE with the ad-
Creep behavior and stress relaxation reflect the dimensional stabi- dition of biochar in terms of flexural properties. Improvement of both
lity of a composite, and also can be used to characterize the mechanical flexural strength and flexural modulus was achieved on the whole. The
properties of a composite. As shown in Fig. 8, the creep behavior and flexural strength and modulus of HDPE were 14.14 MPa and 0.88 GPa,
stress relaxation of neat HDPE and biochar added composites were respectively whereas those of B10 were 16.37 MPa and 0.91 GPa, re-
characterized using creep compliance and stress relaxation modulus in spectively. And this increment of flexural properties was reached due to
this study. Both creep behavior and stress relaxation of all samples the unique characteristics of biochar. Having a high surface area
showed three main stages suggesting the properties of typical thermo- (297.36 m2/g) and a special porous structure should be responsible for
plastic polymers. Interestingly, all the biochar added composites the improvement of flexural properties. Das et al. (Das et al., 2016a)
showed lower creep compliance than that of HDPE, and the creep reported that at high temperatures, HPDE, which was in a fluid state,
compliance continuously decreased with an increase of the biochar could fill the pores of the biochar under the extrusion of the screw, and
amount in the HDPE composites. It can be stated that the addition of a physical/mechanical interlocking was resulted after cooling. Since
biochar improved the creep resistance of HDPE. This can be attributed this physical/mechanical interlocking can transfer stress efficiently, it
to the improvement of dimensional stability, and the addition of rigid was the main contribution for the better flexural properties. It can also

6
Q. Zhang, et al. Industrial Crops & Products 146 (2020) 112185

tensile strength among the samples was obtained. In addition, the


tensile modulus of HDPE, B10, B20, B30, B40, B50, B60 and B70
samples were 1.33, 1.36, 1.47, 1.69, 1.70, 1.87, 2.03 and 0.98 GPa,
respectively. Also, B60 illustrated the maximum tensile modulus with a
significant increase by 52.63 %, from 1.33 GPa for the neat HDPE to
2.03 GPa. Li et al. (Li et al., 2018) have summarized that better tensile
strength of a composite can be obtained by homogenous dispersion of
filler particles in a polymer matrix, which could give an explanation for
better tensile strength of B40 and B50. Also, the stress transfer me-
chanism caused by interfacial bonding dictates the tensile strength of a
composite (Fu et al., 2008). Thus, the betterment of the tensile strength
in biochar added composites was strongly dependent on the afore-
mentioned special interface bonding. The increase of tensile modulus in
biochar added composites suggests another advantage in mechanics,
and this improvement can be attributed to the rigidity of biochars. The
incorporation of rigid biochars into the HDPE matrix reduced the de-
formability and mobility of HDPE macromolecules (Siebert and Wilker,
2019), which improved the stiffness of biochar added composites.
Moreover, biochar obtained at 600 °C carbonization showed low hy-
drophilicity in this study, and this was confirmed by the contact angle
in Fig. S2. Similar polarities resulting in good affinity could also be
responsible for the improvement of tensile properties of biochar added
composites. Nevertheless, the tensile properties of B70 decreased ob-
viously. As can be known from Fig. 1 that B70 was severely deformed.
Limited HDPE matrix cannot afford excessive biochar particles and the
plasticity of the composites was reduced. Excessive biochar caused the
agglomeration, resulting in the fact that the interfacial bonding strength
of B70 was weakened, stress propagation was obstructed, and the ten-
sile properties of B70 were decreased (Nasser et al., 2019).

3.10. Water absorption

In addition to the above mechanical properties, water absorption is


also an important factor affecting the dimensional stability of a com-
Fig. 10. Tensile properties of HDPE and composites samples: (a) tensile
posite. The results of water absorption for HDPE and biochar added
strength (b) tensile modulus.
composites in 28 days at room temperature are presented in Fig. 11.
According to the figure, it is expected that the water absorption of all
be observed from Fig. 9 that flexural strength and modulus of samples the samples increased with the immersion time indicating the damage
B50 and B40 reached a maximum of 34.95 MPa and 1.79 GPa, respec- of dimensional stability by long-time immersion. Besides, HDPE showed
tively. As stated earlier in Table S1, the high SiO2 content (∼89.5 wt%) lower water absorption below 0.2 %, this can be attributed to the mi-
in the ash of biochar was beneficial to improve the mechanical prop- crocracks produced in the manufacturing process. Upon closer ob-
erties of the composites. Bulota et al. obtained a similar result in PLA/ servation in Fig. 11, it can be observed that the water absorption in-
algae composites (Bulota and Budtova, 2015). On the other hand, this creased for the biochar added composites as the biochar loading
particular good increase can also be attributed to the good dispersion of increased, and there was a positive correlation between water absorp-
biochar in HDPE as shown in Fig. S3. Many documents (Goud et al., tion and biochar loading. This result exhibited the worse water-re-
2019; Oliveira et al., 2018) confirmed that the particle dispersion of sistance of biochar added composites than that of neat HDPE with the
fillers in the matrix was important for the flexural strength of a com-
posite. Also, the flexural properties of B60 and B70 confirmed this
theory exactly. From B40 and B50 to B60 and B70, the flexural prop-
erties of biochar added composites showed a significant reduction. The
most important reason for this decrease was the poor particle dispersion
in the matrix. Less HDPE matrix cannot effectively load more biochar
particles, resulting in serious particle agglomeration (Aliotta et al.,
2019) and a compromise of flexural properties.

3.9. Tensile properties

Fig. 10 compares the tensile properties including tensile strength


and tensile modulus of HDPE and biochar added composite samples.
Determining the tensile properties of a composite dictates its applica-
tions, and an overall improvement can be observed from Fig. 10 with
the addition of biochar. The tensile strength of HDPE, B10, B20, B30,
B40, B50, B60 and B70 samples were 23.54, 23.21, 24.88, 25.20, 29.05,
26.25, 24.90 and 19.02 MPa, respectively. When the B40 biochar
loading reached 40 wt%, the tensile strength increased by 23.41 %, Fig. 11. Variation of water absorption in 28 days for HDPE and composites
from 23.54 MPa for the neat HDPE to 29.05 MPa, and the maximum samples.

7
Q. Zhang, et al. Industrial Crops & Products 146 (2020) 112185

and interesting results were obtained in this study. The addition of


biochar did not change the crystal planes of HDPE and no chemical
reaction occurred between HDPE and the biochar. Biochar caused the
early onset of exothermic crystallization of biochar added composites
by DSC and improvement of thermal properties in biochar added
composites was achieved in terms of the amount of residue produced by
TGA. Moreover, the addition of biochar improved the mechanical
properties of HDPE. Best flexural strength of 34.95 MPa in B50, the
flexural modulus of 1.79 GPa in B40, the tensile strength of 29.05 MPa
in B40, and the tensile modulus of 2.03 GPa were obtained, which in-
dicated that high biochar loading filled composites with good me-
chanical properties were achieved. Besides, good stiffness, elasticity,
creep resistance and anti-stress relaxation ability were also obtained in
high biochar added composites. Although the increase of biochar
loadings decreased the water-resistance of the composites, the compo-
Fig. 12. Limited oxygen index of HDPE and composites samples. sites with highly filled biochars suggested better flame retardant. The
results presented here indicated that maximizing the utilization of
agricultural wastes to prepare green products with excellent properties
addition of biochar. The water absorption of a composite is dependent
can be achieved by highly loaded biochar composites.
on some factors such as fine pores, lumens, and the gaps in the interface
(Zabihzadeh, 2010). On the one hand, some interfacial voids between
CRediT authorship contribution statement
biochar and HDPE were produced during the composite making pro-
cess, and water was absorbed by these interfacial voids (Guo et al.,
Qingfa Zhang: Conceptualization, Methodology, Formal analysis,
2019). On the other hand, the porous structure of biochar should also
Investigation, Writing - original draft. Donghong Zhang: Methodology,
be responsible for the poor water resistance, although biochar showed
Writing - review & editing. Hang Xu: Methodology, Writing - review &
low hydrophilicity because of the decrease of hydroxyl groups under
editing. Wenyu Lu: Methodology, Writing - review & editing. Xiajin
high temperatures. Additionally, there was a significant difference in
Ren: Methodology, Writing - review & editing. Hongzhen Cai:
water absorption between B70 and other samples, and B70 showed the
Conceptualization, Funding acquisition, Writing - review & editing,
worst water resistance. This result can be attributed to the agglom-
Supervision. Hanwu Lei: Conceptualization, Funding acquisition,
eration of biochar in HDPE and many biochar particles agglomerated
Writing - review & editing, Supervision. Erguang Huo: Validation,
together, which increased the contact area with water and resulted in
Writing - review & editing. Yunfeng Zhao: Validation, Writing - review
higher water absorption.
& editing. Moriko Qian: Validation, Writing - review & editing. Xiaona
Lin: Validation, Writing - review & editing. Elmar M. Villota: Writing -
3.11. Flame retardancy review & editing. Wendy Mateo: Writing - review & editing.

Since the flammability of a composite dictates application in out- Declaration of Competing Interest
door, limited oxygen index measurements were performed to char-
acterize flame retardancy of HDPE and biochar added composites in The authors declare that they have no known competing financial
this study. Limited oxygen index refers to the volume ratio of oxygen in interests or personal relationships that could have appeared to influ-
the mixture of oxygen and nitrogen., The greater the limited oxygen ence the work reported in this paper.
index, the lower the flammability and the better the flame retardant
effect. As shown in Fig. 12, HDPE shows a lower limited oxygen index Acknowledgments
of 18.21 % indicating its poor flame retardant, and this result was
consistent with the report of Jiang et al. (Jiang et al., 2018). It should The authors appreciate that this work was financially supported by
be noted that the limited oxygen index increased for the biochar added the Natural Science Foundation of Shandong Province of China (No.
composites as the biochar loading increased, which indicates that an ZR2019MEE036); the Agriculture and Food Research Initiative
increase in the flame retardancy was obtained by adding biochar to Competitive Grant (No. 2016-67021-24533, and 2014-38502-22598)
HDPE. As stated earlier, the addition of biochar improved the thermal from the National Institute of Food and Agriculture, United States
properties of HDPE due to its high thermal stability. Thus, it can be Department of Agriculture; the National Key Research and
stated that the improvement of flame retardancy in HDPE was obtained Development Program of China (No. 2018YFD1101001); the National
by the addition of biochar, and this can be attributed to that the biochar Natural Science Fund of China (No. 51806129); the distinguished ex-
with high thermal stability hindered the transport of heat between the pert of Taishan scholars Shandong province project and the higher
heat source and HPDE (Das et al., 2016a). On the other hand, many education superior discipline team training program of Shandong pro-
documents have reported that the preparation of biochar was accom- vince.
panied by the production of a large number of metal oxides and in-
organic substances in the process of carbonization (Chen et al., 2019; Appendix A. Supplementary data
Wei et al., 2018). These nonflammable metal oxides and inorganic
substances acting as flame retardants can effectively slow down the Supplementary material related to this article can be found, in the
combustion of HDPE and improve the flame retardancy of it. Besides, online version, at doi:https://doi.org/10.1016/j.indcrop.2020.112185.
the greatest limited oxygen index of 25.06 % was obtained in B70 in-
dicating the potential of biochar in flame retardant of composites. References

4. Conclusions Aliotta, L., Cinelli, P., Coltelli, M.B., Lazzeri, A., 2019. Rigid filler toughening in PLA-
Calcium Carbonate composites: effect of particle surface treatment and matrix plas-
ticization. Eur. Polym. J. 113, 78–88.
In order to maximize the utilization of agricultural wastes, the effect Arnandha, Y., Satyarno, I., Awaludin, A., Irawati, I.S., Prasetya, Y., Prayitno, D.A.,
of high biochar loadings on biochar added composites was investigated

8
Q. Zhang, et al. Industrial Crops & Products 146 (2020) 112185

Winata, D.C., Satrio, M.H., Amalia, A., 2017. Physical and mechanical properties of acid (PLA) using bamboo charcoal particles. Compos. Part B Eng. 81, 14–25.
WPC board from sengon sawdust and recycled HDPE plastic. Procedia Eng. 171, Jiang, D., Pan, M., Cai, X., Zhao, Y., 2018. Flame retardancy of rice straw-polyethylene
695–704. composites affected by in situ polymerization of ammonium polyphosphate/silica.
Arrigo, R., Jagdale, P., Bartoli, M., Tagliaferro, A., Malucelli, G., 2019. Structure–property Compos. Part A Appl. Sci. Manuf. 109, 1–9.
relationships in polyethylene-based composites filled with biochar derived from Li, S., Li, X., Chen, C., Wang, H., Deng, Q., Gong, M., Li, D., 2016. Development of
waste coffee grounds. Polymers 11, 1336. electrically conductive nano bamboo charcoal/ultra-high molecular weight poly-
Bajwa, D.S., Adhikari, S., Shojaeiarani, J., Bajwa, S.G., Pandey, P., Shanmugam, S.R., ethylene composites with a segregated network. Compos. Sci. Technol. 132, 31–37.
2019. Characterization of bio-carbon and ligno-cellulosic fiber reinforced bio-com- Li, S., Huang, A., Chen, Y.-J., Li, D., Turng, L.-S., 2018. Highly filled biochar/ultra-high
posites with compatibilizer. Constr. Build. Mater. 204, 193–202. molecular weight polyethylene/linear low density polyethylene composites for high-
Behazin, E., Misra, M., Mohanty, A.K., 2017. Sustainable biocarbon from pyrolyzed performance electromagnetic interference shielding. Compos. Part B Eng. 153,
perennial grasses and their effects on impact modified polypropylene biocomposites. 277–284.
Compos. Part B Eng. 118, 116–124. Lim, S.L., Lee, L.H., Wu, T.Y., 2016. Sustainability of using composting and vermi-
Bulota, M., Budtova, T., 2015. PLA/algae composites: morphology and mechanical composting technologies for organic solid waste biotransformation: recent overview,
properties. Compos. Part A Appl. Sci. Manuf. 73, 109–115. greenhouse gases emissions and economic analysis. J. Clean. Prod. 111, 262–278.
Chen, C., Yan, X., Xu, Y., Yoza, B.A., Wang, X., Kou, Y., Ye, H., Wang, Q., Li, Q.X., 2019. Lin, X., Kong, L., Cai, H., Zhang, Q., Bi, D., Yi, W., 2019. Effects of alkali and alkaline
Activated petroleum waste sludge biochar for efficient catalytic ozonation of refinery earth metals on the co-pyrolysis of cellulose and high density polyethylene using TGA
wastewater. Sci. Total Environ. 651, 2631–2640. and Py-GC/MS. Fuel Process. Technol. 191, 71–78.
Chow, C.-F., Wong, W.-L., Chan, C.-W., Chan, C.-S., 2018. Converting inert plastic waste Mashouf Roudsari, G., Mohanty, A.K., Misra, M., 2017. A statistical approach to develop
into energetic materials: a study on the light-accelerated decomposition of plastic biocomposites from epoxy resin, poly (furfuryl alcohol), poly (propylene carbonate),
waste with the Fenton reaction. Waste Manag. 75, 174–180. and biochar. J. Appl. Polym. Sci. 134, 45307.
Das, O., Sarmah, A.K., Bhattacharyya, D., 2015. A sustainable and resilient approach Mohanty, A.K., Vivekanandhan, S., Pin, J.-M., Misra, M., 2018. Composites from re-
through biochar addition in wood polymer composites. Sci. Total Environ. 512, newable and sustainable resources: challenges and innovations. Science 362,
326–336. 536–542.
Das, O., Bhattacharyya, D., Hui, D., Lau, K.-T., 2016a. Mechanical and flammability Nan, N., DeVallance, D.B., Xie, X., Wang, J., 2016. The effect of bio-carbon addition on
characterisations of biochar/polypropylene biocomposites. Compos. Part B Eng. 106, the electrical, mechanical, and thermal properties of polyvinyl alcohol/biochar
120–128. composites. J. Compos. Mater. 50, 1161–1168.
Das, O., Bhattacharyya, D., Sarmah, A.K., 2016b. Sustainable eco–composites obtained Nasser, J., Lin, J., Steinke, K., Sodano, H.A., 2019. Enhanced interfacial strength of
from waste derived biochar: a consideration in performance properties, production aramid fiber reinforced composites through adsorbed aramid nanofiber coatings.
costs, and environmental impact. J. Clean. Prod. 129, 159–168. Compos. Sci. Technol. 174, 125–133.
Das, O., Sarmah, A.K., Zujovic, Z., Bhattacharyya, D., 2016c. Characterisation of waste Oliveira, L.Á., Santos, J.C., Panzera, T.H., Freire, R.T., Vieira, L.M., Scarpa, F., 2018.
derived biochar added biocomposites: chemical and thermal modifications. Sci. Total Evaluation of hybrid-short-coir-fibre-reinforced composites via full factorial design.
Environ. 550, 133–142. Compos. Struct. 202, 313–323.
Das, O., Kim, N.K., Kalamkarov, A.L., Sarmah, A.K., Bhattacharyya, D., 2017. Biochar to Pandey, P., Bajwa, S.G., Bajwa, D.S., Englund, K., 2017. Performance of UV weathered
the rescue: balancing the fire performance and mechanical properties of poly- HDPE composites containing hull fiber from DDGS and corn grain. Ind. Crops Prod.
propylene composites. Polym. Degrad. Stab. 144, 485–496. 107, 409–419.
Das, O., Kim, N.K., Hedenqvist, M.S., Lin, R.J., Sarmah, A.K., Bhattacharyya, D., 2018. An Pillai, K.V., Renneckar, S., 2016. Dynamic mechanical analysis of layer-by-layer cellulose
attempt to find a suitable biomass for biochar-based polypropylene biocomposites. nanocomposites. Ind. Crops Prod. 93, 267–275.
Environ. Manage. 62, 403–413. Poulose, A.M., Elnour, A.Y., Anis, A., Shaikh, H., Al-Zahrani, S., George, J., Al-Wabel,
Das, O., Hedenqvist, M.S., Johansson, E., Olsson, R.T., Loho, T.A., Capezza, A.J., Raman, M.I., Usman, A.R., Ok, Y.S., Tsang, D.C., 2018. Date palm biochar-polymer compo-
R.S., Holder, S., 2019. An all-gluten biocomposite: comparisons with carbon black sites: an investigation of electrical, mechanical, thermal and rheological character-
and pine char composites. Compos. Part A Appl. Sci. Manuf. 120, 42–48. istics. Sci. Total Environ. 619, 311–318.
Davis, A.M., Hanzly, L.E., DeButts, B.L., Barone, J.R., 2019. Characterization of dimen- Santhiago, M., Garcia, P.S., Strauss, M., 2018. Bio-based nanostructured carbons toward
sional stability in flax fiber reinforced polypropylene composites. Polym. Compos. 40, sustainable technologies. Curr. Opin. Green Sustain. Chem. 12, 22–26.
132–140. Siebert, H.M., Wilker, J.J., 2019. Deriving commercial level adhesive performance from a
De Bhowmick, G., Sarmah, A.K., Sen, R., 2018. Production and characterization of a value bio-based mussel mimetic polymer. ACS Sustain. Chem. Eng. 7, 13315–13323.
added biochar mix using seaweed, rice husk and pine sawdust: a parametric study. J. Song, B., Chen, M., Zhao, L., Qiu, H., Cao, X., 2019. Physicochemical property and col-
Clean. Prod. 200, 641–656. loidal stability of micron-and nano-particle biochar derived from a variety of feed-
Demirbas, A., 2011. Waste management, waste resource facilities and waste conversion stock sources. Sci. Total Environ. 661, 685–695.
processes. Energy Convers. Manage. 52, 1280–1287. Turku, I., Keskisaari, A., Kärki, T., Puurtinen, A., Marttila, P., 2017. Characterization of
Fa, W., Wang, J., Ge, S., Chao, C., 2019. Performance of photo-degradation and thermo- wood plastic composites manufactured from recycled plastic blends. Compos. Struct.
degradation of polyethylene with photo-catalysts and thermo-oxidant additives. 161, 469–476.
Polym. Bull. 1–16. Wang, X., Peng, S., Chen, H., Yu, X., Zhao, X., 2019. Mechanical properties, rheological
Ferreira, G.F., Pierozzi, M., Fingolo, A.C., da Silva, W.P., Strauss, M., 2019. Tuning su- behaviors, and phase morphologies of high-toughness PLA/PBAT blends by in-situ
garcane bagasse biochar into a potential carbon black substitute for polyethylene reactive compatibilization. Compos. Part B Eng., 107028.
composites. J. Polym. Environ. 27, 1735–1745. Wei, D., Li, B., Huang, H., Luo, L., Zhang, J., Yang, Y., Guo, J., Tang, L., Zeng, G., Zhou, Y.,
Fu, S.-Y., Feng, X.-Q., Lauke, B., Mai, Y.-W., 2008. Effects of particle size, particle/matrix 2018. Biochar-based functional materials in the purification of agricultural waste-
interface adhesion and particle loading on mechanical properties of particulate–po- water: fabrication, application and future research needs. Chemosphere 197,
lymer composites. Compos. Part B Eng. 39, 933–961. 165–180.
Giorcelli, M., Khan, A., Pugno, N.M., Rosso, C., Tagliaferro, A., 2019. Biochar as a cheap Zabihzadeh, S.M., 2010. Water uptake and flexural properties of natural filler/HDPE
and environmental friendly filler able to improve polymer mechanical properties. composites. BioResources 5, 316 -232.
Biomass Bioenergy 120, 219–223. Zeng, S., Shen, M., Yang, L., Xue, Y., Lu, F., Chen, S., 2018. Self-assembled montmor-
Goud, V., Alagirusamy, R., Das, A., Kalyanasundaram, D., 2019. Influence of various illonite–carbon nanotube for epoxy composites with superior mechanical and thermal
forms of polypropylene matrix (fiber, powder and film states) on the flexural strength properties. Compos. Sci. Technol. 162, 131–139.
of carbon-polypropylene composites. Compos. Part B Eng. 166, 56–64. Zhang, Q., Yi, W., Li, Z., Wang, L., Cai, H., 2018. Mechanical properties of rice husk
Guo, G., Finkenstadt, V.L., Nimmagadda, Y., 2019. Mechanical properties and water biochar reinforced high density polyethylene composites. Polymers 10, 286.
absorption behavior of injection-molded wood fiber/carbon fiber high-density poly- Zhang, Q., Khan, M.U., Lin, X., Cai, H., Lei, H., 2019a. Temperature varied biochar as a
ethylene hybrid composites. Adv. Compos. Hybrid Mater. 1–11. reinforcing filler for high-density polyethylene composites. Compos. Part B Eng. 175,
Hahn, A., Gerdts, G., Völker, C., Niebühr, V., 2019. Using FTIRS as pre-screening method 107151.
for detection of microplastic in bulk sediment samples. Sci. Total Environ. 689, Zhang, Q., Li, Y., Cai, H., Lin, X., Yi, W., Zhang, J., 2019b. Properties comparison of high
341–346. density polyethylene composites filled with three kinds of shell fibers. Results Phys.
Ho, M.-p., Lau, K.-t., Wang, H., Hui, D., 2015. Improvement on the properties of polylactic 12, 1542–1546.

You might also like