Open access peer-reviewed chapter

Towards Traditional Carbon Fillers: Biochar-Based Reinforced Plastic

Written By

Mattia Bartoli, Mauro Giorcelli, Pravin Jagdale and Massimo Rovere

Submitted: 30 September 2019 Reviewed: 02 March 2020 Published: 28 March 2020

DOI: 10.5772/intechopen.91962

From the Edited Volume

Fillers

Edited by Emmanuel Flores Huicochea

Chapter metrics overview

1,036 Chapter Downloads

View Full Metrics

Abstract

The global market of carbon-reinforced plastic represents one of the largest economic platforms. This sector is dominated by carbon black (CB) produced from traditional oil industry. Recently, high technological fillers such as carbon fibres or nanostructured carbon (i.e. carbon nanotubes, graphene, graphene oxide) fillers have tried to exploit their potential but without economic success. So, in this chapter we are going to analyse the use of an unconventional carbon filler called biochar. Biochar is the solid residue of pyrolysis and can be a solid and sustainable replacement for traditional and expensive fillers. In this chapter, we will provide overview of the last advancement in the use of biochar as filler for the production of reinforced plastics.

Keywords

  • biochar
  • composites
  • sustainable production
  • carbon materials

1. Introduction

Carbon-based materials are a very well-established commodity generally used in materials science [1]. Nowadays, many commodities makes use of carbon fibres as they become an unavoidable asset for the global market [2]. Carbon fibres represent the most diffuse high-tech carbon materials, but carbon black played the main role in these materials. Carbon black (CB) harvests a great global carbon revenue due to its use for the production of plenty of composites but mainly for tyres [3]. Over the years, high-cost carbon materials such as carbon nanotubes (CNTs) and graphene-like materials have gained the attention of the scientific community with their amazing conductivity and optic and mechanical properties [4, 5]. Despite the expected revolution, nanosized allotropic carbon form did not have much progress in the research area. In a very optimistic report, Segal [6] dreamed that the world was ready for the industrial-scale production of graphene, but after a decade, single-layer graphene is still sold at 200 €/cm2, while graphene oxide costs 100,000 €/kg [7]. On the other hand, CB is sold for around 1 €/kg [8]. New-generation high-tech carbon materials (i.e. CNT, graphene and graphene oxide) have not yet fulfilled the promise for a new carbon era. While the industry waits for large-scale commercialization of high-quality affordable carbon allotropes, new materials have been considered through engineered carbon for profitable business. In recent years, a new material has emerged as the most promising for the integration of carbon production with waste management [9, 10, 11]. This material is biochar, the solid residue from pyrolytic conversion of biomass. Biomass waste stream is one of the most abundant worldwide, and it is generally disposed through incineration. This presents both an environmental issue and an economic loss due to the transformation of a high-quality material into heat. Accordingly, a more profitable advantage was found in their thermal conversion for the production of biofuels [12, 13], chemicals [14] and other materials [15]. Conversion of biomass into liquid fuels is challenging due to the high oxygen content compared with traditional oil-derived products (i.e. gasoline, virgin nafta, diesel). On the contrary, biochar production is a process full of opportunities with the emergence of carbonaceous material from both lignocellulosic and non-lignocellulosic biomasses. This bioderived carbon is used in many applications [16] due to its properties and low cost attested at around 1–2 €/kg [17, 18, 19]. Actually, biochar has found a large-scale application for soil health improvement [20, 21, 22] and as solid fuel with a heating content of around 40 kJ/mol [23]. Nonetheless, these applications are limited and unable to exploit the full potentiality of biochar due to their easy tunability with simple process adjustment [24].

In this book chapter, we report an overview of the composite applications of biochar to prove its feasibility as a replacement for traditional carbon materials and as a solid competitor with high-tech reinforced plastics.

Advertisement

2. Biochar: production ways

Biochar is produced through thermochemical routes such as torrefaction, pyrolysis and hydrothermal carbonisation and as residue of gasification.

Torrefaction is a low-temperature thermal conversion used to densify the biomasses for energy purposes [25]. The process temperature is in the range from 200–350°C, and the conversion requires long residence and processing times. Torrefaction is characterised by biochar and biochar-like yields [26]. The carbon percentage of solid residue is generally around 50–60 wt.% [27], but it can reach 72–80 wt.% using microwave process with the addition of microwave absorbers as reported in several studies [28, 29, 30, 31]. Microwave use leads to the drastic reduction of process timescale from hours to minutes.

Pyrolysis is a high-temperature thermochemical conversion which induces the cracking of polymers with the formation of low-molecular-weight compounds in an oxygen-free or oxygen-poor atmosphere [12, 32]. Pyrolysis is run using different heating technologies [33] and apparatus design [34, 35, 36, 37] at a temperature range from 450–700°C [38] with huge variations in product fraction yields. Pyrolysis of biomasses was deeply studied, and the main mechanisms can be rationalised in a few different steps. The first is the release of moisture from the feedstock, increasing the surface area and improving the pore structure, which favours a quick release of volatiles and minimizes char-catalysed secondary cracking. Lignocellulosic biomass behaviour during pyrolysis could be rationalised through the behaviour of the main components as cellulose, hemicellulose and lignin. The pyrolysis of cellulose takes place between 430 and 470°C and hemicellulose between 470 and 600°C while lignin between 600 and 800°C. During this process other reactions such as dehydration of the sample, pyrolysis of the volatiles present, formation of levoglucosan from cellulose [39] and formation of substituted aromatic rings from lignin [40] take place together with the formation of carbon.

Hydrothermal carbonisation is a thermal cracking used to produce crude-like oil and hydrochar under moderate temperature and high pressure [41] using aqueous solvent [42], nonaqueous solvent [43, 44] or subcritical/critical media [45].

Finally, gasification is the conversion of biomass into combustible gas by heating in air [46], pure oxygen or steam [47] at temperatures higher than 800°C with or without a catalyst [48]. Products from gasification are a mixture of carbon monoxide, carbon dioxide, methane, hydrogen and water vapour. Biochar is not the main product of gasification, but it is characterised by a simultaneous high carbon and ash content.

In summary, Figure 1 shows the main stage of biomass conversion.

Figure 1.

Conversion of lignocellulosic biomasses to carbon structures.

Furthermore, it is relevant to notice that leaves, stems, bark and roots are different lignin/cellulose ratio and mechanical properties. The same is true for different species.

Some of these differences could be retained into biochar and induce appreciable properties.

Also, the graphitic domains formed during pyrolytic treatment could undergo a stacking rearrangement leading to a graphitization of biochar with the increasing temperature.

Advertisement

3. Biochar-based reinforced plastics

Nowadays, reinforced plastic materials are one of the largest global markets in the polymer sector with an expected global revenue of up to 130 M$/year in 2024 as summarised in Figure 2.

Figure 2.

Reinforced plastic worldwide revenue with a prediction for the year 2024 [49].

Carbon-containing reinforced plastic is one of the most relevant materials with an annual production of up to 150 kton/y in 2018 [50]. As clearly reported in Figure 3, around 80% of the total carbon-containing reinforced plastic is represented by polymer host materials, 49% of which comes from thermoset resin and 30% from thermoplastic polymers. Among them, carbon fibre-reinforced epoxy polymers are the larger amount. This is due to their numerous applications in all-capital high-tech sectors ranging from aeronautics and aerospace industries [51] to automotive industries [52]. In this global scenario, biochar plays a minor role even if it could be used, and it is going to consolidate itself as a trustworthy commodity with its production flexibility and property tunability [53].

Figure 3.

World’s carbon-based reinforced plastic production in 2018 [50].

Therefore, the main uses of biochar in both thermoset and thermoplastic matrices are overviewed.

Carbonaceous-reinforced thermoset resins are the most commonly used materials dispersed in plenty of different polymer hosts [54, 55, 56]. Epoxy resins are the most deeply applied and used around the world. Consequently, the replacement of carbon fibres and carbon black, carbon soot and anthracites with biochar has gained a great interest. Khan et al. [57] described the mechanical and dielectric properties of high-temperature-annealed maple-derived biochar dispersed into a two-component epoxy resin. Biochar filler was used in concentration ranging from 0.5 wt.% to 20 wt.%. The authors clearly showed the improvement of mechanical properties using filler loading of up to 4 wt.%. Regarding electric properties, Khan and co-workers found that a low loading of multiwalled CNTs induced the same effects of a 20 wt.% loading of biochar. Recently, Bartoli et al. [58] described the relationship between the biochar morphology and related composite mechanical properties using a biochar loading of 2 wt.%. The authors achieved a 40% increment of maximum elongation using a rhizomatous grass-derived biochar and Young’s modulus increment using a wheat straw as a source for biochar production. The authors suggested that smooth surface could induce an improved mobility inside the epoxy matrix, while highly porous and channelled surfaces do not. Additionally, they suggested that the dispersion methodology adopted based on ultrasonication, summarised in Figure 4, reduced the size of biochar particles with a direct relation with the original morphology of the very same particles.

Figure 4.

Ultrasonication methodology for the effect dispersion of biochar inside the epoxy resin host.

Furthermore, pyrolytic temperature is the main and critical parameter for tuning biochar properties with the goal of improving resin properties. The interactions between epoxy resin and biochar particles. Bartoli et al. [59] studied the effect of the heating rate and maximum pyrolytic temperature on biochar. Furthermore, cellulose templates could be used for the production of biochar fibres and balls using selected precursors. As a matter of fact, biochar produced from wasted cotton fibres could be recovered as carbon fibre shape showing a property enhancement of epoxy resin host matrix [60, 61], while the one produced from cellulose nanocrystals could be recovered as micrometrics ball or nanometric needles [62].

Authors showed the complex relationship of produced biochar with related containing epoxy composites. Sample prepared at different temperature and using different heating rate increment or the Young’s modulus or the toughness of the reinforced plastics. Interestingly, the biochar produced at very high temperature of up to 1000°C generally induced a high increment of elongation probably due to the unpacking of the aromatic ring of epoxy host.

Similarly, Giorcelli et al. [63] proved the effectiveness of maple tree-derived biochar produced at 600°C and 1000°C, observing a drastic improvement of maximum elongation compared with neat resin.

Temperature also affected the electrical properties of biochar and biochar-containing composites. High thermal annealed biochar could represent a solid choice for the production of conductive epoxy composites. Giorcelli et al. [64] described that highly graphitic biochar induced better performances during DC electrical conductivity measurements. Temperature treatment and related graphitization processes lead to an improved ability of these materials to shield microwave radiation with similar outputs with respect to multiwalled CNTs [65] even under thin-film shape [66].

The other huge field of composite materials is represented by thermoplastic reinforced plastics. In this very same field, polyolefin represents the greater amount of worldwide production. Among them, biochar-containing polyethylene was studied by Arrigo et al. [67] using an exhausting coffee-derived biochar produced at 700°C. The authors described the rheological and thermal properties of biochar-related composites with a filler loading up to 7.5 wt. %, showing a decrement of the dynamics of polymer chains in the host matrix related to the confinement of the polymer chains on the biochar surface. Additionally, the well-embedded biochar particles improved the thermo-oxidative stability of polyethylene composites produced. Zhang et al. [68] studied the temperature influence on biochar production from poplar and its use as filler for high-density polyethylene. Curiously, the microcrystalline structure of the polymer was not affected by the presence of biochar according to the thermal data collected. A different trend was reported for the mechanical properties that were appreciable different in the comparison between neat and biochar-loaded poly(ethylene) with an improvement of flexural strength and a decrement of the impact strength. Zhang et al. [69] valourized agricultural waste streams through pyrolysis, and the resulting biochar was used as filler for ultra-high-density poly(ethylene). The authors observed improvement in their mechanical properties and improvement in the flame retardancy of the high filler loading materials. Similar results were achieved by Sundarakannan et al. [70] using biochar derived from cashew nuts. Also Li et al. [71] investigated the high loads of biochar in ultra-high-molecular-weight poly(ethylene), achieving a remarkable electromagnetic interference shielding properties using an 80 wt.% of bamboo biochar pyrolysed at 1100°C. This material showed a very high conductivity of up to 107.6 S/m. Furthermore, Bajwa et al. [72] described the use of biochar for the production of a composite blend based on high-density poly(ethylene), poly(lactic acid) and wood flour with superior thermal stability.

The other largely used polyolefins is poly(propylene). Poly(propylene) is widely applied for the realisation of biochar-based composites due its workability. The main application of poly(propylene)-related composites is in the automotive sector. Tadele et al. [73] published an interesting comparative study on life cycle assessment of biochar used in automotive, showing the feasibility of its use. Das et al. [74] reached the same conclusions about the economic feasibility of the use of biochar instead of traditional carbonaceous fillers. The authors showed the appreciable cost reduction of a biochar-containing composites achieving the same properties of carbon black-based ones due to the sensible reduction of compatibilizer down to a maximum of 3 wt.%. The affordability of the cost of biochar was the core of the research proposed by Behazin et al. [75]. In this study, a pyrolysed perennial cane was used as filler of a polymer blend based on poly(propylene)/poly(octene-ethylene). The produced composite contained a filler loading ranging from 10 to 20 wt.% and showed the very strong interactions between polymer matrix and biochar particles. The most detailed and comprehensive set of studies about poly(propylene) and biochar interactions was conducted by Bhattacharyya research group and his co-workers as attested by many papers [76, 77, 78]. During this pluri-annual research, the authors investigated the kind and magnitude of interactions between filler and poly(propylene). They conclude that the addition of several types of biochars lead to a general improvement of the mechanical and thermal properties of related poly(propylene) composites together with the induction of flame retardancy properties. Additionally, Elnour et al. [79] studied the relationship between biochar properties and related poly(propylene) composites showing an increment of stiffness together with unaffected tensile strength. In the same period, Poulose et al. [80] mixed date palm-derived biochar with poly(propylene) showing the negligible effect of biochar on the storage modulus in a range of concentration of up to 15 wt.%. Poly(ethylene) and poly(propylene) are not the only polyolefins used for the production of carbon-based composites. Other largely used polyolefin matrices were poly(vinyl alcohol) and its derivatives [81, 82] and poly(acrylonitrile) [83]. Both of those two polymeric hosts are used for the realisation of piezo sensors due to their elastic properties.

Furthermore, polyesters were used for the realisation of carbon-based reinforced materials. As an example, polyamides were impregnated with biochar as described by Ogunsona et al. [84]. The authors mixed nylon 6 with the biochar produced from the pyrolysis of Miscanthus canes. The biochar used was produced using a process temperature ranging from 500–900°C. The different temperatures used affected the output of the composites with a beneficial effect on only the high-temperature-treated biochar and a detrimental effect on the others. In 2019, Sheng et al. [85] modified bamboo biochar through the addition of silyl groups on the particle surface for the production of poly(lactic acid) composites. Surface functionalization showing an appreciable enhancement of maximum elongation of up to 93% was compared with neat polymer matrix.

Recently, biochar was used for the realisation of biopolymer composites based on polysaccharides such as cellulose [86], starch [87] or gluten [88] under the vision of blue and green economy for a total bio- and sustainable productive line.

Advertisement

4. Conclusions

In this chapter, we provided overview of the most recent applications of biochar in the field of polymer composite production with a focus on more useful and unusual ones. We also described in detail the possibility of using biochars as a sound replacement for traditional fillers in both thermoset and thermoplastic composite materials. The researches herein described the feasibility of biochar used in different industrial sectors as a solid alternative to traditional and nanostructured materials. The adaptive nature of biochar presents a very strong point of advantage for spreading its use across the field of materials science.

References

  1. 1. Burchell TD. Carbon Materials for Advanced Technologies. Amsterdam, The Netherlands: Elsevier; 1999
  2. 2. Holmes M. Global carbon fibre market remains on upward trend. Reinforced Plastics. 2014;58:38-45
  3. 3. I.C.B. Association. What Is Carbon Black?; 2019. Availabe online: http://www.carbon-black.org/ [Accessed: 17 December 2019]
  4. 4. Endo M. Carbon nanotube research: Past and future. Japanese Journal of Applied Physics. 2012;51:040001
  5. 5. Singh V, Joung D, Zhai L, Das S, Khondaker SI, Seal S. Graphene based materials: Past, present and future. Progress in Materials Science. 2011;56:1178-1271
  6. 6. Segal M. Selling graphene by the ton. Nature Nanotechnology. 2009;4:612
  7. 7. Welcome To Cheap Tubes. Availabe online: https://www.cheaptubes.com/ [Accessed: 9 October 2019]
  8. 8. Alibaba.com. Market Price for Carbon Black n550. Alibaba.com; 2019
  9. 9. Bezama A, Agamuthu P. Addressing the Big Issues in Waste Management. London, England: SAGE Publications Sage UK; 2019
  10. 10. Mui EL, Ko DC, McKay G. Production of active carbons from waste tyres—A review. Carbon. 2004;42:2789-2805
  11. 11. Couth R, Trois C. Carbon emissions reduction strategies in Africa from improved waste management: A review. Waste Management. 2010;30:2336-2346
  12. 12. Bridgwater AV. Review of fast pyrolysis of biomass and product upgrading. Biomass and Bioenergy. 2012;38:68-94
  13. 13. Damartzis T, Zabaniotou A. Thermochemical conversion of biomass to second generation biofuels through integrated process design—A review. Renewable and Sustainable Energy Reviews. 2011;15:366-378
  14. 14. Amen-Chen C, Pakdel H, Roy C. Production of monomeric phenols by thermochemical conversion of biomass: A review. Bioresource Technology. 2001;79:277-299
  15. 15. Paz-Ferreiro J, Nieto A, Méndez A, Askeland MPJ, Gascó G. Biochar from biosolids pyrolysis: A review. International Journal of Environmental Research and Public Health. 2018;15:956
  16. 16. Qian K, Kumar A, Zhang H, Bellmer D, Huhnke R. Recent advances in utilization of biochar. Renewable and Sustainable Energy Reviews. 2015;42:1055-1064
  17. 17. Biochar: Prospects of Commercialization. Availabe online: https://farm-energy.extension.org/ [Accessed: 14 November 2019]
  18. 18. Vochozka M, Maroušková A, Váchal J, Straková J. Biochar pricing hampers biochar farming. Clean Technologies and Environmental Policy. 2016;18:1225-1231
  19. 19. Maroušek J. Significant breakthrough in biochar cost reduction. Clean Technologies and Environmental Policy. 2014;16:1821-1825
  20. 20. Ali S, Rizwan M, Qayyum MF, Ok YS, Ibrahim M, Riaz M, et al. Biochar soil amendment on alleviation of drought and salt stress in plants: A critical review. Environmental Science and Pollution Research. 2017;24:12700-12712
  21. 21. Maroušek J, Strunecký O, Stehel V. Biochar farming: Defining economically perspective applications. Clean Technologies and Environmental Policy. 2019;21:1-7
  22. 22. Maroušek J, Kolář L, Vochozka M, Stehel V, Maroušková A. Biochar reduces nitrate level in red beet. Environmental Science and Pollution Research. 2018;25:18200-18203
  23. 23. Abdullah H, Wu H. Biochar as a fuel: 1. Properties and grindability of biochars produced from the pyrolysis of mallee wood under slow-heating conditions. Energy & Fuels. 2009;23:4174-4181
  24. 24. Al-Wabel MI, Al-Omran A, El-Naggar AH, Nadeem M, Usman ARA. Pyrolysis temperature induced changes in characteristics and chemical composition of biochar produced from conocarpus wastes. Bioresource Technology. 2013;131:374-379
  25. 25. Chen W-H, Peng J, Bi XT. A state-of-the-art review of biomass torrefaction, densification and applications. Renewable and Sustainable Energy Reviews. 2015;44:847-866
  26. 26. Gronnow MJ, Budarin VL, Mašek O, Crombie KN, Brownsort PA, Shuttleworth PS, et al. Torrefaction/biochar production by microwave and conventional slow pyrolysis—Comparison of energy properties. GCB Bioenergy. 2013;5:144-152
  27. 27. Bach Q-V, Chen W-H, Chu Y-S, Skreiberg Ø. Predictions of biochar yield and elemental composition during torrefaction of forest residues. Bioresource Technology. 2016;215:239-246
  28. 28. Bartoli M, Rosi L, Giovannelli A, Frediani P, Frediani M. Characterization of bio-oil and bio-char produced by low-temperature microwave-assisted pyrolysis of olive pruning residue using various absorbers. Waste Management & Research. 2020;38(2):213-225
  29. 29. Bartoli M, Rosi L, Giovannelli A, Frediani P, Frediani M. Bio-oil from residues of short rotation coppice of poplar using a microwave assisted pyrolysis. Journal of Analytical and Applied Pyrolysis. 2016;119:224-232
  30. 30. Bartoli M, Rosi L, Giovannelli A, Frediani P, Frediani M. Pyrolysis of α-cellulose in a microwave multimode batch reactor. Journal of Analytical and Applied Pyrolysis. 2016;120:284-296
  31. 31. Bartoli M, Rosi L, Giovannelli A, Frediani P, Frediani M. Production of bio-oils and bio-char from Arundo donax through microwave assisted pyrolysis in a multimode batch reactor. Journal of Analytical and Applied Pyrolysis. 2016;122:479-489
  32. 32. Wampler TP. Applied Pyrolysis Handbook. Boca Raton, FL, USA: CRC Press; 2006
  33. 33. Bridgwater A, Peacocke G. Fast pyrolysis processes for biomass. Renewable and Sustainable Energy Reviews. 2000;4:1-73
  34. 34. Ragucci R, Giudicianni P, Cavaliere A. Cellulose slow pyrolysis products in a pressurized steam flow reactor. Fuel. 2013;107:122-130
  35. 35. Tsai W, Lee M, Chang Y. Fast pyrolysis of rice straw, sugarcane bagasse and coconut shell in an induction-heating reactor. Journal of Analytical and Applied Pyrolysis. 2006;76:230-237
  36. 36. Karaduman A, Şimşek EH, Çiçek B, Bilgesü AY. Flash pyrolysis of polystyrene wastes in a free-fall reactor under vacuum. Journal of Analytical and Applied Pyrolysis. 2001;60:179-186
  37. 37. Li D, Briens C, Berruti F. Improved lignin pyrolysis for phenolics production in a bubbling bed reactor—Effect of bed materials. Bioresource Technology. 2015;189:7-14
  38. 38. Horne PA, Williams PT. Influence of temperature on the products from the flash pyrolysis of biomass. Fuel. 1996;75:1051-1059
  39. 39. Zhang X, Yang W, Dong C. Levoglucosan formation mechanisms during cellulose pyrolysis. Journal of Analytical and Applied Pyrolysis. 2013;104:19-27
  40. 40. Kotake T, Kawamoto H, Saka S. Mechanisms for the formation of monomers and oligomers during the pyrolysis of a softwood lignin. Journal of Analytical and Applied Pyrolysis. 2014;105:309-316
  41. 41. Behrendt F, Neubauer Y, Oevermann M, Wilmes B, Zobel N. Direct liquefaction of biomass. Chemical Engineering & Technology. 2008;31:667-677
  42. 42. Akhtar J, Amin NAS. A review on process conditions for optimum bio-oil yield in hydrothermal liquefaction of biomass. Renewable and Sustainable Energy Reviews. 2011;15:1615-1624
  43. 43. Liu Z, Zhang F-S. Effects of various solvents on the liquefaction of biomass to produce fuels and chemical feedstocks. Energy Conversion and Management. 2008;49:3498-3504
  44. 44. Zou S, Wu Y, Yang M, Li C, Tong J. Thermochemical catalytic liquefaction of the marine microalgae Dunaliella tertiolecta and characterization of bio-oils. Energy & Fuels. 2009;23:3753-3758
  45. 45. Peterson AA, Vogel F, Lachance RP, Fröling M, Antal MJ Jr, Tester JW. Thermochemical biofuel production in hydrothermal media: A review of sub-and supercritical water technologies. Energy & Environmental Science. 2008;1:32-65
  46. 46. Guizani C, Sanz FE, Salvador S. Influence of temperature and particle size on the single and mixed atmosphere gasification of biomass char with H2O and CO2. Fuel Processing Technology. 2015;134:175-188
  47. 47. Barisano D, Canneto G, Nanna F, Alvino E, Pinto G, Villone A, et al. Steam/oxygen biomass gasification at pilot scale in an internally circulating bubbling fluidized bed reactor. Fuel Processing Technology. 2016;141:74-81
  48. 48. Cheah S, Jablonski WS, Olstad JL, Carpenter DL, Barthelemy KD, Robichaud DJ, et al. Effects of thermal pretreatment and catalyst on biomass gasification efficiency and syngas composition. Green Chemistry. 2016
  49. 49. Grand View Research. Market Value of Composite Materials Worldwide from 2015 to 2024 (in Billion U.S. Dollars). Grand View Research, statista.com; 2019
  50. 50. Sauer M, Kühnel M, Witten E. Composites Market Report 2018—Market Developments, Trends, Outlook and Challenges. Germany: AVK-TV Industrievereinigung Verstrkte Kunststoffe Carbon Composite; 2018
  51. 51. Chand S. Review carbon fibers for composites. Journal of Materials Science. 2000;35:1303-1313
  52. 52. Adam H. Carbon fibre in automotive applications. Materials & Design. 1997;18:349-355
  53. 53. Downie A, Crosky A, Munroe P. Physical properties of biochar. In: Biochar for Environmental Management: Science and Technology. London, UK: MPG Book; 2009. pp. 13-32
  54. 54. Bartoli M, Rosi L, Frediani M. Chapter 22—Synthesis and applications of unsaturated polyester composites. In: Thomas S, Hosur M, Chirayil CJ, editors. Unsaturated Polyester Resins. Amsterdam, The Netherlands: Elsevier; 2019. pp. 579-598
  55. 55. Rana S, Alagirusamy R, Joshi M. A review on carbon epoxy nanocomposites. Journal of Reinforced Plastics and Composites. 2009;28:461-487
  56. 56. Kim S, Pechar TW, Marand E. Poly (imide siloxane) and carbon nanotube mixed matrix membranes for gas separation. Desalination. 2006;192:330-339
  57. 57. Khan A, Savi P, Quaranta S, Rovere M, Giorcelli M, Tagliaferro A, et al. Low-cost carbon fillers to improve mechanical properties and conductivity of epoxy composites. Polymers. 2017;9:642
  58. 58. Bartoli M, Giorcelli M, Rosso C, Rovere M, Jagdale P, Tagliaferro A. Influence of commercial biochar fillers on brittleness/ductility of epoxy resin composites. Applied Sciences. 2019;9:3109
  59. 59. Bartoli M, Nasir MA, Jagdale P, Passaglia E, Spiniello R, Rosso C, et al. Influence of pyrolytic thermal history on olive pruning biochar and related epoxy composites mechanical properties. Journal of Composite Materials. 2019:11
  60. 60. Jagdale P, Koumoulos EP, Cannavaro I, Khan A, Castellino M, Dragatogiannis DA, et al. Towards green carbon fibre manufacturing from waste cotton: A microstructural and physical property investigation. Manufacturing Review. 2017;4:10
  61. 61. Khan A, Jagdale P, Castellino M, Rovere M, Jehangir Q , Mandracci P, et al. Innovative functionalized carbon fibers from waste: How to enhance polymer composites properties. Composites Part B: Engineering. 2018;139:31-39
  62. 62. Bartoli M, Giorcelli M, Jagdale P, Rovere M, Tagliaferro A, Chae M, et al. Shape tunability of carbonized cellulose nanocrystals. SN Applied Sciences. 2019;1:1661-1676
  63. 63. Giorcelli M, Khan A, Pugno NM, Rosso C, Tagliaferro A. Biochar as a cheap and environmental friendly filler able to improve polymer mechanical properties. Biomass and Bioenergy. 2019;120:219-223
  64. 64. Giorcelli M, Savi P, Khan A, Tagliaferro A. Analysis of biochar with different pyrolysis temperatures used as filler in epoxy resin composites. Biomass and Bioenergy. 2019;122:466-471
  65. 65. Giorcelli M, Khan A, Tagliaferro A, Savi P, Berruti F. Microwave characterization of polymer composite based on biochar: A comparison of composite behaviour for biochar and MWCNTs. In: 2016 IEEE International Nanoelectronics Conference (INEC). Chengdu, China: IEEE; 9-11 May 2016. pp. 1-2
  66. 66. Quaranta S, Savi P, Giorcelli M, Khan AA, Tagliaferro A, Jia CQ. Biochar-polymer composites and thin films: Characterizations and applications. In: 2016 IEEE 2nd International Forum on Research and Technologies for Society and Industry Leveraging a Better Tomorrow (RTSI); 2016. pp. 1-4
  67. 67. Arrigo R, Jagdale P, Bartoli M, Tagliaferro A, Malucelli G. Structure–property relationships in polyethylene-based composites filled with biochar derived from waste coffee grounds. Polymers. 2019;11:1336
  68. 68. Zhang Q , Khan MU, Lin X, Cai H, Lei H. Temperature varied biochar as a reinforcing filler for high-density polyethylene composites. Composites Part B: Engineering. 2019;175:107151
  69. 69. Zhang Q , Zhang D, Xu H, Lu W, Ren X, Cai H, et al. Biochar filled high-density polyethylene composites with excellent properties: Towards maximizing the utilization of agricultural wastes. Industrial Crops and Products. 2020;146:112185
  70. 70. Sundarakannan R, Arumugaprabu V, Manikandan V, Vigneshwaran S. Mechanical property analysis of biochar derived from cashew nut shell waste reinforced polymer matrix. Materials Research Express. 2020;6:125349
  71. 71. Li S, Huang A, Chen Y-J, Li D, Turng L-S. Highly filled biochar/ultra-high molecular weight polyethylene/linear low density polyethylene composites for high-performance electromagnetic interference shielding. Composites Part B: Engineering. 2018;153:277-284
  72. 72. Bajwa DS, Adhikari S, Shojaeiarani J, Bajwa SG, Pandey P, Shanmugam SR. Characterization of bio-carbon and ligno-cellulosic fiber reinforced bio-composites with compatibilizer. Construction and Building Materials. 2019;204:193-202
  73. 73. Tadele D, Roy P, Defersha F, Misra M, Mohanty AK. A comparative life-cycle assessment of talc-and biochar-reinforced composites for lightweight automotive parts. Clean Technologies and Environmental Policy. 2020:1-11
  74. 74. Das O, Bhattacharyya D, Sarmah AK. Sustainable eco–composites obtained from waste derived biochar: A consideration in performance properties, production costs, and environmental impact. Journal of Cleaner Production. 2016;129:159-168
  75. 75. Behazin E, Misra M, Mohanty AK. Sustainable biocarbon from pyrolyzed perennial grasses and their effects on impact modified polypropylene biocomposites. Composites Part B: Engineering. 2017;118:116-124
  76. 76. Das O, Kim NK, Kalamkarov AL, Sarmah AK, Bhattacharyya D. Biochar to the rescue: Balancing the fire performance and mechanical properties of polypropylene composites. Polymer Degradation and Stability. 2017;144:485-496
  77. 77. Das O, Sarmah AK, Bhattacharyya D. Biocomposites from waste derived biochars: Mechanical, thermal, chemical, and morphological properties. Waste Management. 2016;49:560-570
  78. 78. Das O, Bhattacharyya D, Hui D, Lau K-T. Mechanical and flammability characterisations of biochar/polypropylene biocomposites. Composites Part B: Engineering. 2016;106:120-128
  79. 79. Elnour AY, Alghyamah AA, Shaikh HM, Poulose AM, Al-Zahrani SM, Anis A, et al. Effect of pyrolysis temperature on biochar microstructural evolution, physicochemical characteristics, and its influence on biochar/polypropylene composites. Applied Sciences. 2019;9:1149
  80. 80. Poulose AM, Elnour AY, Anis A, Shaikh H, Al-Zahrani SM, George J, et al. Date palm biochar-polymer composites: An investigation of electrical, mechanical, thermal and rheological characteristics. Science of the Total Environment. 2018;619-620:311-318
  81. 81. Nan N, DeVallance DB, Xie X, Wang J. The effect of bio-carbon addition on the electrical, mechanical, and thermal properties of polyvinyl alcohol/biochar composites. Journal of Composite Materials. 2016;50:1161-1168
  82. 82. Nan N, DeVallance DB. Development of poly (vinyl alcohol)/wood-derived biochar composites for use in pressure sensor applications. Journal of Materials Science. 2017;52:8247-8257
  83. 83. Nan W, Zhao Y, Ding Y, Shende AR, Fong H, Shende RV. Mechanically flexible electrospun carbon nanofiber mats derived from biochar and polyacrylonitrile. Materials Letters. 2017;205:206-210
  84. 84. Ogunsona EO, Misra M, Mohanty AK. Impact of interfacial adhesion on the microstructure and property variations of biocarbons reinforced nylon 6 biocomposites. Composites Part A: Applied Science and Manufacturing. 2017;98:32-44
  85. 85. Sheng K, Zhang S, Qian S, Fontanillo Lopez CA. High-toughness PLA/bamboo cellulose nanowhiskers bionanocomposite strengthened with silylated ultrafine bamboo-char. Composites Part B: Engineering. 2019;165:174-182
  86. 86. Zhang Q , Lei H, Cai H, Han X, Lin X, Qian M, et al. Improvement on the properties of microcrystalline cellulose/polylactic acid composites by using activated biochar. Journal of Cleaner Production. 2020;252:119898
  87. 87. George J, Azad LB, Poulose AM, An Y, Sarmah AK. Nano-mechanical behaviour of biochar-starch polymer composite: Investigation through advanced dynamic atomic force microscopy. Composites Part A: Applied Science and Manufacturing. 2019;124:105486
  88. 88. Das O, Hedenqvist MS, Johansson E, Olsson RT, Loho TA, Capezza AJ, et al. An all-gluten biocomposite: Comparisons with carbon black and pine char composites. Composites Part A: Applied Science and Manufacturing. 2019;120:42-48

Written By

Mattia Bartoli, Mauro Giorcelli, Pravin Jagdale and Massimo Rovere

Submitted: 30 September 2019 Reviewed: 02 March 2020 Published: 28 March 2020