Open access peer-reviewed chapter

A Mini Review of Biochar Synthesis, Characterization, and Related Standardization and Legislation

Written By

Nor Adilla Rashidi and Suzana Yusup

Submitted: 24 December 2019 Reviewed: 22 April 2020 Published: 22 May 2020

DOI: 10.5772/intechopen.92621

From the Edited Volume

Applications of Biochar for Environmental Safety

Edited by Ahmed A. Abdelhafez and Mohammed H. H. Abbas

Chapter metrics overview

1,179 Chapter Downloads

View Full Metrics

Abstract

The abundance of biomass in Malaysia creates an avenue for growth of bio-economic sector through the research and development (R&D) activities on the biochar production. Biochar that is described as a carbonaceous material derived from the thermochemical process at temperature of usually lower than 700°C is promising due to its applicability in wider range of applications, such as in soil amendment (fertilizer) and as a low-cost adsorbent for the pollution remediation, apart from minimizing the solid waste disposal problems. Therefore, this chapter discusses the current trends on various production techniques of biochar from both the lignocellulosic (plantation based waste materials) and non-lignocellulosic sources, as well as the physiochemical characteristics of the resulting biochar. In addition, overview of the biochar industry in Malaysia is presented in this chapter. Lastly, recap of standardization and legislation particularly related to the biochar utilization as a soil amendment agent is included to grasp readers’ attention prior to the large scale applications.

Keywords

  • biochar
  • biomass
  • environmental standardization and legislation
  • pyrolysis
  • soil amendment
  • waste management

1. Introduction

Biochar, which is a subset of carbon-rich and black powder, is generally defined as a porous solid that is produced from biomass via pyrolysis process and in the absence of oxygen (O2) [1]. Nevertheless, based on literatures, there are various definitions of the biochar [2, 3, 4]; accordingly, Sohi et al. [5] reported that the term of biochar remains ill-defined. Thus, the International Biochar Initiative (IBI) standardized the biochar “as a solid material obtained from the thermochemical conversion of biomass in O2-limited environments.” While the production route of biochar and charcoal is similar where both materials are derived from the carbonaceous feedstock through the pyrolysis process [6], but the distinct features that can distinguish these two materials lies in their starting material and end application. Biochar that contains high porosity, high nutrient content, and water-storage-capability is applied for soil amelioration or an adsorbent, whereas charcoal that is usually derived from the petroleum-based feedstock is used for heat generation (energy/fuel) purposes [3, 7]. In a nutshell, Mesa et al. [2] reported that the term biochar is not applicable for the charred materials used as a solid fuel, and to exclude the black carbon produced from non-renewable resources such as coal and petroleum. Besides, Abdelhafez et al. [8] reported that biochar contains lower ash compounds as compared to charcoal, due to an incomplete carbonization process. Further, due to wider application of biochar in both agronomic sector as well as in environmental management, Verheijen et al. [9] reported that the global market of biochar is rapidly growing, with the global market price is estimated around $80–13,480/oven dried metric ton (ODMT). In addition, Hersh et al. [10] reported that the global biochar market is projected to increase up to $3.14 billion by 2025, and expand at an average rate of 13.1% annually [11]. Due to the growing interest of the biochar production and application, number of scientific publications related to the biochar is gradually increasing (as presented in Figure 1), where most of these publications (since 2016) are from Republic of China, USA, Australia, South Korea, and India. Herein, this chapter aims to highlight the recent advancement of the biochar production from various processing techniques, as well as an overview on the biochar standardization (quality standard) and legislation, particularly for its application as soil amendment agent.

Figure 1.

Number of biochar-related publications from 2009 to 2019 from Web of Science (WOS) and SCOPUS database.

So far, research work on the biochar-related field in Malaysia is extensive in local universities and research institutes, where Universiti Putra Malaysia (UPM) is the leading organization in the biochar research. Being a pioneer in biochar research, UPM researchers in collaboration with Nasmech Technology has successfully built the first large scale biochar production plant within the region (as shown in Figure 2) in January 2010 [12, 13], where the carbonator is capable to accommodate up to 20 tons of different types of waste materials daily for the biochar production. Hypothetically, opportunities of biochar industry in Malaysia can be attributed to lower labor cost, low or no cost incurred of biomass, large agricultural industry, as well as fast-growing biomass. In fact, Ozturk et al. [14] reported that Malaysia produces about 168 million tons of biomass annually. Nevertheless, Kong et al. [15] reported that the main challenge in biochar production in Malaysia is due to the physiochemical nature of biomass (particularly oil palm biomass) itself; where wet biomass will result in transportation problems from the source to production sites, thus need an additional drying process apart from normal pre-treatments such as chopping, shredding, and grinding stages. Consequently, this will increase both production cost and equipment’s capital investment. Besides, difficulty in gaining a long-term contract basis between the biomass suppliers, producers, investors, and potential end users, is one of the major barriers in the biochar production in Malaysia [15, 16]. Due to these problems and the lack of key players along the value chain, biochar’s production is rather costly, accordingly, Tang et al. [17] reported that commercialization of biochar in Malaysia is relatively new and still at an early stage. Based on literatures, biochar providers in Malaysia include the following: Global Green Synergy Sdn. Bhd., Pakar Go Green Sdn. Bhd., Usaha Strategik Sdn. Bhd., and CH Biotech Sdn. Bhd. In addition, realizing the prominence of the biochar industry toward the socio-environmental economy, Biochar Association Malaysia (BMA) has been established in 2014 with the missions are to promote the biochar production and application in both agricultural and industrial sector, to stimulate publics’ awareness on the role of biochar as a carbon sequester, and as a platform for idea and information exchange in promoting the biochar industry in Malaysia. In addition, to further promote the advancement of biochar industry in Malaysia, key players including researchers, authorities, and business analysts should work closely together.

Figure 2.

Biochar plant in Dengkil, Selangor, Malaysia [18].

Advertisement

2. Production of biochar

Biochar can be produced from various types of biomass which include the lignocellulosic (i.e., bioenergy crop, agricultural waste, forestry residues) and non-lignocellulosic groups (i.e., manure, sewage sludge, microalgae) [19, 20]. To date, agricultural waste is the primary feedstock used for the biochar production, as confirmed in Table 1. Regardless of the different types of feedstock, the biochar’s skeleton is primarily comprised of carbon and ash, where the overall compositions and characteristics of each biochar varied, depending on the types of feedstock and the process conditions. Filiberto et al. [20] reported that the significant difference between the nutrient-rich feedstocks such as animal manure and sewage sludge, compared to the lignin-rich biomass feedstock is that the former materials contain considerably high nutrient and mineral compositions (i.e., nitrogen, phosphorus, potassium, etc.). In context of heavy metal removal application, Zhao et al. [21] reported that the sewage sludge biochar that has higher mineral contents (161 g/kg) compared to corn biochar (28.6 g/kg) and poplar wood biochar (19.5 g/kg) contributes to higher heavy metal removals from wastewater (sewage sludge > corn > poplar wood), thus implies the importance of the mineral compositions in heavy metal adsorption process. Likewise, for the soil amendment application (in terms of element supplementation and liming effect), Zhang et al. [22] also agreed that the biochar should contain a sufficient mineral composition. Meanwhile in context of the process technologies, biochar can be produced from four thermochemical routes that include pyrolysis, torrefaction, hydrothermal carbonization, as well as gasification, [23]; which is thoroughly described in the following subsections.

2.1 Pyrolysis

By definition, pyrolysis is the thermal conversion process conducted in absence of O2; producing biochar, condensable liquid (i.e., bio-oil), and non-condensable gas (i.e., syngas). The yield distribution depends on the type of pyrolysis process—slow, fast, and flash pyrolysis; where it differs in terms of reaction temperature, heating rate, and holding time (as summarized in Table 2).

BiomassProcess conditionsFindingsRef.
Temp (°C)time (min)Yield (%)Capacity
Heavy metals removal (i.e., cadmium, copper, lead, zinc, etc.)
Cocoa pod500120n/a69.9 mg/g[24]
EFB61512825.4915.18 mg/g[25]
EFB300180n/a85 mg/g[26]
Sludgen/a60n/a19 mg/g[27]
Sludge4009064.248.8 mg/g[28]
Color/dyes removal (i.e., methylene blue, malachite green)
Cassava stem50012011.9440.5 mg/g[29]
Coconut frond800240n/a126.58 mg/g[30]
Palm shell700 W253348 mg/g[31]
Seaweed80090n/a512.67 mg/g[32]
Sugarcane bagasse600120n/a99.47%[33]
Phenolic compounds removal
EFB50080.27n/a7.38%[34]
Gas/vapor adsorption (i.e., CO2, mercury, sulfur dioxide)
Coconut pith9006027.766067.49 μg/g[35]
Coconut pith7006031.4210 mmol/g[36]
Sludge4058854.259.75 mg/g[37]
Wood sawdust65060n/a18 mg/g[38]
Soil-based application (herbicides/pesticides removal, fertilizer)
EFB30060n/a4.497[39]
Rice husk300180n/a4.742[39]
Palm shell700 W2533450 g[31]

Table 1.

Summary of recent biochar production in Malaysia from local biomass and the corresponding optimum conditions.

EFB, empty fruit bunches.

ConditionsSlow pyrolysisFast pyrolysisFlash pyrolysis
Temperature (°C)300–700550–1000800–1100
Heating rate (°C/sec)0.1–110–200>1000
Vapor residence time (sec)450–5500.5–10<0.5
Particle size (mm)5–50<1<0.2
Yield (wt. %)Biochar352012
Bio-oil305075
Syngas353013

Table 2.

Process conditions for slow (conventional), fast, and flash pyrolysis and product distribution [40, 41].

Bold value refers to the highest product yield of slow pyrolysis, fast pyrolysis and flash pyrolysis. In summary, for the slow pyrolysis, the bold value is for biochar, while for fast and flash pyrolysis, the bold value is for the bio-oil. In other words, the slow pyrolysis favors the biochar production, and both fast and flash pyrolysis targets the bio-oil.

Referring to Table 2, the ideal route for the biochar production is through slow pyrolysis, also known as conventional carbonization, as compared to fast or flash pyrolysis that targets bio-oil production. Recently, Yuan et al. [42] confirmed that walnut shell biochar obtained through slow pyrolysis process has greater biochar yield as compared to the fast pyrolysis, irrespective of reaction temperature, thus it confirms the effectiveness of the slow pyrolysis mechanism toward the biochar production. Furthermore, slow pyrolysis for the biochar production is promising due to lower capital investment as compared to fast pyrolysis scheme ($132 vs. $200 million) [43]. Basically, Daful et al. [44] reported that biochar from slow pyrolysis route refers to primary and secondary char, where the mechanism of the process is simplified in Eqs. (1)(3) [45]. The pre-pyrolysis reaction [Eq. (1)] involves the water elimination and evaporation from the biomass structure. During the primary reaction, devolatilization process including the dehydration, decarboxylation, and dehydrogenation occurs. Then upon the completion of primary decomposition, the secondary reaction (at high temperature) that refers to cracking of heavy organic compounds as well as repolymerization ensues, producing a stable and carbon-dense solid product (i.e., biochar) and non-condensable syngas such as methylene (CH2), methane (CH4), carbon monoxide (CO), and carbon dioxide (CO2) [45, 46, 47, 48].

Biomasswater+unreacted residuePrepyrolysisE1
Unreacted residuevolatiles+gases+charPrimary reactionE2
Charvolatiles+gases+charSecondary reactionE3

2.2 Torrefaction

Torrefaction or known as a mild pyrolysis refers to the thermochemical process at temperature of 200–300°C at atmospheric pressure and inert atmosphere, heating rate of ≤50°C/min, with residence time of 30 min to 2 h [44, 49]. Nevertheless, it has been reported that the torrefaction process is not a promising technique for the biochar production, regardless of higher product yield (70–80 wt. %), since the torrefied biomass still contains a significant fraction of volatile components from the raw biomass, and the physiochemical properties are in between raw biomass and biochar [44, 50]. For example, oxygen to carbon (O/C) ratio of the torrefied biomass which is >0.4 contradicts with the European Biochar Certification (EBC) of biochar [44]. Therefore, this torrefaction process is often being applied as a pre-treatment process for moisture removal, biomass densification, and to improve the biomass properties. Besides, while the torrefaction process alone cannot be used for biochar production, combination of torrefaction pretreatment and pyrolysis is feasible for the exceptional biochar production (in terms of yield) in addition to the physiochemical characteristics (i.e., surface area) [51, 52, 53, 54].

2.3 Hydrothermal carbonization

Opposite to the slow pyrolysis and torrefaction process that is normally carried out under dry atmosphere, hydrothermal carbonization can also be referred as wet pyrolysis or wet torrefaction; since this process is performed in a biomass-water solution at temperature of 180–250°C at high pressure (subcritical condition) for several hours [50, 55, 56, 57]. Similar to pyrolysis, this hydrothermal carbonization produces 50–80 wt. % solid char (termed as hydrochar), bio-oil and water mixture (5–20 wt. %), and synthetic gas that is mainly CO2 (2–5 wt. %) [58]. The great interest in this hydrothermal technology for the biochar production is that it can avoid the preliminary energy-intensive drying process that is usually required for the conventional pyrolysis, and thus it will minimize the operational costs. Besides, Oktaviananda et al. [59] agreed that such process is convenient for the biomass having >50 wt. % moisture content. On top of that, it has been reported that the energy requirement for hydrothermal carbonization and pyrolysis process for 1 kg of feedstock of 80% moisture content is 2.5 and 3.20 MJ, respectively [60]. Moreover, this hydrothermal technology offers the lowest reaction temperature as compared to other thermochemical conversion techniques. During the process, water (H2O) acts as a solvent, reactant, catalyst, and as a medium for both mass and energy transfer [61], where it will facilitate the hydrolysis, dehydration, decarboxylation and depolymerization process [62]. Besides, at temperature of 200–280°C, H2O that possesses similar behavior to mild acid and mild base at the same time results in an acceleration of biomass decomposition [61, 63]. Specifically, Libra et al. [64] reported that during the hydrothermal carbonization, hemicellulose decomposes at temperature of 180–200°C, lignin decomposition takes place at 180–220°C, whereas cellulose decomposition occurs at 220°C. However, most often, the hydrochar cannot be described as biochar since the reaction temperature is too low, low carbon contents, as well as an intolerable O/C and hydrogen to carbon (H/C) ratio [65, 66]. Yet, recent work shows that integration of this hydrothermal carbonization with pyrolysis process positively contributes toward the high-quality biochar production and can stabilizes the heavy metal in solid products [67]. For example, by referring to the experimental findings by Olszewski et al. [68], the preliminary hydrothermal treatment of brewery spent grains (that contains 70–90 wt. % moisture) prior to the pyrolysis process produces biochar with greater product yield and carbon contents as well as reduced ash compositions; where the corresponding value is varied, subjected to the intensity of the hydrothermal carbonization process. Likewise, Garlapalli et al. [69] confirmed that the carbon compositions of biochar from the combined hydrothermal and pyrolysis process (at 260 and 800°C, respectively) increases to 82 wt. % compared to standalone hydrothermal process, where the carbon contents is merely 70 wt. %. Moreover, such combined processes also show an improvement of the surface area (63.48 m2/g vs. 2.93 m2/g). In overall, the upgrading of hydrochar is crucial since the hydrochar that possesses low surface area (<30 m2/g), low porosity, and presence of noxious chemicals (i.e., furan, furfural, and phenolic compounds) limits its application in soil amelioration [69].

2.4 Gasification

The gasification process takes place at the temperature range of 600-1200°C, heating rate of 50–100°C/min, with vapor residence time of 10–20 s. Unlike the pyrolysis, gasification process is carried out in the presence of O2 (including O2, air, steam, CO2, or mixture of the gases) and primarily used for the syngas production (i.e., CO, CO2, CH4, hydrogen [H2]) instead of the biochar production. Due to this, the biochar yield is minimal (<10 wt. %) [44, 56]. With regards to this limitation, there are limited research works on the feasibility of biochar from the gasification process especially for soil amendment purpose [70]. In addition, Wang and Wang [71] reported that the charred product from the gasification process do not satisfy the biochar’s definition; in addition to presence of hazardous polycyclic aromatic hydrocarbons (PAHs) as well as alkaline and alkaline heavy metals within the structure [55, 56].

Advertisement

3. Biochar’s characterization, standardization, and legislations

The detailed characterization of biochar prior to any applications is significant in order to determine the relationship between nature and operating conditions with the physiochemical properties of biochar, to evaluate the suitability of biochar in desired target application, and to examine the presence of contaminants and eco-toxicology properties [72]. The overall characterization techniques that have been applied for biochar are summarized in Table 3.

CharacterizationDetailed analysis
Physical property
  • Surface area, pore volume and size (N2 gas sorption)

  • Particle size distribution (Laser sizing)

  • Density (Mercury porosity, Pycnometer)

Chemical property
  • pH (pH meter)

  • Electrical conductivity (Conductivity meter)

  • Cation exchange capacity (Ion chromatography)

  • Biochar compositions (CHNS, EDS, XPS)

  • Metallic/ash contents (XRD, ICP, XRF)

  • Proximate analysis (Muffle furnace, TGA)

  • Surface functionality (FTIR, Raman)

  • Surface acidity/alkalinity (Boehm titration)

  • Surface aromaticity (13C NMR, Raman spectroscopy)

Surface structure & morphology
  • SEM/FESEM

  • TEM

  • Crystallinity (XRD, Raman)

Stability behavior
  • TGA-DSC

Table 3.

Summary of biochar’s detailed characterization [19, 46, 64, 71, 72, 73, 74].

Given that the biochar’s characteristics is mainly influenced by various parameters such as feedstocks’ type, technology (i.e., process type, reactor configuration), and process condition (i.e., temperature, heating rate, residence time, pressure, carrier gas); the corresponding properties of biochar are widely varied. Therefore, the standardization of biochar prior to applications is significant as their performance can be generalized and predicted [64, 75]. To date, the biochar standards have been established by the International Biochar Initiative (IBI-BS), European Biochar Foundation (European Biochar Certificate, EBC); as well as the British Biochar Foundation (Biochar Quality Mandate, BQM) [76, 77, 78]. Referring to Verheijen et al. [9], the common objectives of these certifications are to provide the quality and safety indicator for biochar utilization as a soil amendment agent, to promote the biochar’s industrial growth and commercialization, as well as for future legislative or regulations. Besides, development of such certifications assists in improving the confidence level of consumers and regulators of the biochar’s safe application [79]. Thereby, the parameters and their corresponding threshold values in each biochar certificate are tabulated in Table 4.

PropertyIBI-BSEBCBQM
BasicPremiumStandardHigh gr.
Organic C (wt. %)≥10≥50≥10
H:C molar ratio≤0.7≤0.7≤0.7
O:C molar ratio≤0.4
Moisture≥30≥20
Total ash (wt. %)
ConductivityOptional
Liming equiv.
pH
Particle size distr.
Surface areaOptional
Water holding capacityOptional
Volatile matter (%)Optional
Germination testPass/failOptional
Macro-nutrients (wt. %)
Total N
Total P, K, Mg, CaOptional✓ (Total P & K)
Organic pollutants (mg/kg)
PAH
(US EPA 16)
6–300<12<4<20<20
B(a) P toxic equi.≤3
PCB0.2–0.5<0.2<0.5
PCDDs/Fs<17<20<20
Heavy metals (mg/kg)—maximum limit
Arsenic12–10010010
Cadmium1.4–391.51393
Chromium64–1200908010015
Cobalt40–150
Copper63–1500100100150040
Lead70–50015012050060
Mercury1–1711171
Manganesen/a3500
Molybdenum5–207510
Nickel47–600503060010
Selenium2–361005
Zinc200–70004004002800150
Boron
Chlorine
Sodium

Table 4.

Summary of biochar certification based on IBI-BS (Ver. 2.0), EBC (Ver. 4.8), and BQM Ver. 1.0.

Note: symbol refers to the required analysis for biochar (declaration).

However, Gelardi et al. [80] reported that variation between these certifications will led to inconsistencies in both scientific and legislative framework, accordingly, there is an urgent need to come out with a unified regulations that can benefit the communication in academics field and in the biochar market. In addition, it should be noted that these certifications are only applicable for the biochar categorization and their suitability as soil amendment agent, and to exclude the hydrochar [65]. Hence, more data and research work toward the hydrochar characterization and appropriate certificates that enable commercial hydrochar utilization is strongly recommended. In addition, since these certifications are only valid for the biochar usage in soil application, it is recommended to produce a detailed assessment and guideline for the biochar utilization in other environmental applications too [74].

Advertisement

4. Conclusions and future outlook

Biomass valorization to biochar materials has gained a significant attention due to its exceptional characteristics—high surface area, high pore volume, long-term stability, and presence of various surface functionalities, as well as wider potential application including energy and biomaterial development, agronomy sector (i.e., soil amelioration, fertilization), and environment pollution control; among others. Given the slow pyrolysis process is the most promising technique for the biochar production, more research studies on the various types of biomass need to be considered as the biochar field is rather a non-exhaustive subject, in addition to the continuous advancement toward cleaner, simpler, and inexpensive biochar production. In addition, a comprehensive analysis on different types of biomass (including agricultural, aquaculture, forestry, human and animal waste, as well as industrial waste) will result in a complete database; mainly focus on the influence of operating parameters toward the process performance, in terms of reaction rate and underlying mechanism, yield, selectivity, biochar’s characteristics, as well as energy and mass balance; which are useful for practitioners and future researchers. In addition, from the databases, it is practical for ranking the biomass suitability for the biochar production for specific applications, accordingly facilitates a proper planning on biomass utilization in biochar industry. Besides, the recent work on both the biochar production and utilization is limited to the laboratory scale, thus upscaling the research work to a larger scale is necessary in order to determine the practicality. Finally, techno-economic analysis as well as life cycle assessment of the biochar production through various technologies is recommended. Overall, viability of the biochar industrial sector needs to incorporate the social, technical, economic, and environmental aspects to ensure its sustainability.

Advertisement

Acknowledgments

The authors would like to acknowledge the Ministry of Education Malaysia (MOE) for awarding the National Higher Institution Centre of Excellence (HICoE) award to Centre for Biofuel and Biochemical Research, Universiti Teknologi PETRONAS (UTP).

Advertisement

Conflict of interest

The authors declare no conflict of interest.

References

  1. 1. Zhang H, Tu Y-J, Duan Y-P, Liu J, Zhi W, Tang Y, et al. Production of biochar from waste sludge/leaf for fast and efficient removal of diclofenac. Journal of Molecular Liquids. 2020;299:112193
  2. 2. Mesa AC, Spokas KA. Impacts of biochar (black carbon) additions on the sorption and efficacy of herbicides. Herbicides and Environment. 2011;13:315-340
  3. 3. Ahmad M, Rajapaksha AU, Lim JE, Zhang M, Bolan N, Mohan D, et al. Biochar as a sorbent for contaminant management in soil and water: A review. Chemosphere. 2014;99:19-33
  4. 4. Rajapaksha AU, Mohan D, Igalavithana AD, Lee SS, Ok Y. Definitions and fundamentals of biochar. In: Ok YS, Uchimiya SM, Chang SX, Bolan N, editors. Biochar: Production, Characterization, and Applications. New York: CRC Press; 2016. pp. 4-17
  5. 5. Sohi SP, Krull E, Lopez-Capel E, Bol R. A review of biochar and its use and function in soil. In: Sparks DL, editor. Advances in Agronomy. United States: Academic Press; 2010. pp. 47-82
  6. 6. Aller MF. Biochar properties: Transport, fate, and impact. Critical Reviews in Environmental Science and Technology. 2016;46:1183-1296
  7. 7. Akdeniz N. A systematic review of biochar use in animal waste composting. Waste Management. 2019;88:291-300
  8. 8. Abdelhafez AA, Abbas MH, Li J. Biochar: The black diamond for soil sustainability, contamination control and agricultural production. In: Huang WJ, editor. Engineering Applications of Biochar. Croatia: IntechOpen; 2017. pp. 7-27
  9. 9. Verheijen FG, Bastos AC, Schmidt H-P, Jeffery S. Biochar and certification 1. In: Vogt M, editor. Sustainability Certification Schemes in the Agricultural and Natural Resource Sectors: Outcomes for Society and the Environment. London and New York: Routledge; 2019. pp. 113-136
  10. 10. Hersh B, Mirkouei A, Sessions J, Rezaie B, You Y. A review and future directions on enhancing sustainability benefits across food-energy-water systems: The potential role of biochar-derived products. AIMS Environmental Science. 2019;6:379-416
  11. 11. Guo M, Xiao P, Li H. Valorization of agricultural byproducts through conversion to biochar and bio-oil. In: Simpson BK, Aryee AN, Toldrá F, editors. Byproducts from Agriculture and Fisheries: Adding Value for Food, Feed, Pharma, and Fuels. USA: John Wiley & Sons; 2020. pp. 501-522
  12. 12. Abdullah N, Sulaiman F. The oil palm wastes in Malaysia. In: Matovic MD, editor. Biomass Now: Sustainable Growth and Use. Croatia: IntechOpen; 2013. pp. 75-93
  13. 13. Harsono SS, Grundman P, Lau LH, Hansen A, Salleh MAM, Meyer-Aurich A, et al. Energy balances, greenhouse gas emissions and economics of biochar production from palm oil empty fruit bunches. Resources, Conservation and Recycling. 2013;77:108-115
  14. 14. Ozturk M, Saba N, Altay V, Iqbal R, Hakeem KR, Jawaid M, et al. Biomass and bioenergy: An overview of the development potential in Turkey and Malaysia. Renewable and Sustainable Energy Reviews. 2017;79:1285-1302
  15. 15. Kong S-H, Loh S-K, Bachmann RT, Rahim SA, Salimon J. Biochar from oil palm biomass: A review of its potential and challenges. Renewable and Sustainable Energy Reviews. 2014;39:729-739
  16. 16. Norli I, Fazilah A, Pazli IM. Agricultural biomass utilisation as a key driver for Malaysian bioeconomy. In: Dabbert S, Lewandowski I, Weiss J, Pyka A, editors. Knowledge-Driven Developments in the Bioeconomy: Technological and Economic Perspectives. Switzerland: Springer; 2017. p. 141–159
  17. 17. Tang KM, Ibrahim WA, Kadir WR. Towards environmental and economic sustainability via the biomass industry: The Malaysian case study. In: Bruckman VJ, Varol EA, Uzun BB, Liu J, editors. Biochar: A Regional Supply Chain Approach in View of Climate Change Mitigation. United Kingdom: Cambridge University Press; 2016. pp. 162-183
  18. 18. UPM-Nasmech Carbonator Pilot Plant, Dengkil Selangor [Internet]. 2010. Available from: http://biocharmalaysia.blogspot.com/2010/01/carbonator-pilot-plant.html [Accessed: 23 March 2020]
  19. 19. Nartey OD, Zhao B. Biochar preparation, characterization, and adsorptive capacity and its effect on bioavailability of contaminants: An overview. Advances in Materials Science and Engineering. 2014;2014:1-12
  20. 20. Filiberto DM, Gaunt JL. Practicality of biochar additions to enhance soil and crop productivity. Agriculture. 2013:715-725
  21. 21. Zhao J, Shen X-J, Domene X, Alcañiz J-M, Liao X, Palet C. Comparison of biochars derived from different types of feedstock and their potential for heavy metal removal in multiple-metal solutions. Scientific Reports. 2019;9:1-12
  22. 22. Zhang H, Chen C, Gray EM, Boyd SE. Effect of feedstock and pyrolysis temperature on properties of biochar governing end use efficacy. Biomass and Bioenergy. 2017;105:136-146
  23. 23. Bartoli M, Giorcelli M, Jagdale P, Rovere M, Tagliaferro A. A review of non-soil biochar applications. Materials. 2020;13:1-35
  24. 24. Yong SK, Leyom J, Tay CC, Talib SA. Sorption of lead from aqueous system using cocoa pod husk biochar: Kinetic and isotherm studies. International Journal of Engineering & Technology. 2018;7:241-244
  25. 25. Zamani SA, Yunus R, Samsuri A, Salleh M, Asady B. Removal of zinc from aqueous solution by optimized oil palm empty fruit bunches biochar as low cost adsorbent. Bioinorganic Chemistry and Applications. 2017;2017:1-9
  26. 26. Sadegh-Zadeh F, Samsuri AW, Seh-Bardan BJ, Emadi M. The effects of acidic functional groups and particle size of biochar on Cd adsorption from aqueous solutions. Desalination and Water Treatment. 2017;66:309-319
  27. 27. Lee XJ, Lee L, Hiew B, Gan S, Thangalazhy-Gopakumar S. Evaluation of the effectiveness of low cost adsorbents from oil palm wastes for wastewater treatment. Chemical Engineering Transactions. 2017;56:937-942
  28. 28. Goh CL, Sethupathi S, Bashir MJ, Ahmed W. Adsorptive behaviour of palm oil mill sludge biochar pyrolyzed at low temperature for copper and cadmium removal. Journal of Environmental Management. 2019;237:281-288
  29. 29. Zaid M, Jamion N, Omar Q, Yong S. Sorption of malachite green (MG) by cassava stem biochar (CSB) kinetic and isotherm studies. Journal of Fundamental and Applied Sciences. 2017;9:273-287
  30. 30. Mohammad R, Rajoo AT, Mohamad M. Coconut fronds as adsorbent in the removal of malachite green dye. ARPN Journal of Engineering and Applied Sciences. 2017;12:996-1001
  31. 31. Liew RK, Nam WL, Chong MY, Phang XY, Su MH, Yek PNY, et al. Oil palm waste: An abundant and promising feedstock for microwave pyrolysis conversion into good quality biochar with potential multi-applications. Process Safety and Environment Protection. 2018;115:57-69
  32. 32. Ahmed M, Okoye P, Hummadi E, Hameed B. High-performance porous biochar from the pyrolysis of natural and renewable seaweed (Gelidiella acerosa) and its application for the adsorption of methylene blue. Bioresource Technology. 2019;278:159-164
  33. 33. Mohamad M, Mohammad R, May TS, Wei LJ. Removal of malachite green by sugarcane bagasse biochar using response surface methodology. AIP Conference Proceedings. 2019;2068:1-6
  34. 34. Arshad SHM, Ngadi N, Wong S, Amin NS, Razmi FA, Mohamed NB, et al. Optimization of phenol adsorption onto biochar from oil palm empty fruit bunch (EFB). Malaysian Journal of Fundamental and Applied Sciences. 2019;15:1-5
  35. 35. Johari K, Saman N, Song ST, Cheu SC, Kong H, Mat H. Development of coconut pith chars towards high elemental mercury adsorption performance—Effect of pyrolysis temperatures. Chemosphere. 2016;156:56-68
  36. 36. Rahim ARA, Kuanaseaan K, Shehzad N, Rabat NE, Johari K, Mat H. Synthesis and characterization of coconut pith char adsorbents for carbon dioxide capture. Malaysian Journal of Fundamental and Applied Sciences. 2019;15:803-805
  37. 37. Iberahim N, Sethupathi S, Bashir MJ. Optimization of palm oil mill sludge biochar preparation for sulfur dioxide removal. Environmental Science and Pollution Research. 2018;25(26):25702-25714
  38. 38. Ghani W, Azlina W, Da Silva G. Sawdust-derived biochar: Characterization and CO2 adsorption/desorption study. Journal of Applied Sciences. 2014;14:1450-1454
  39. 39. Yavari S, Malakahmad A, Sapari NB, Yavari S. Synthesis optimization of oil palm empty fruit bunch and rice husk biochars for removal of imazapic and imazapyr herbicides. Journal of Environmental Management. 2017;193:201-210
  40. 40. Uddin M, Techato K, Taweekun J, Rahman MM, Rasul M, Mahlia T, et al. An overview of recent developments in biomass pyrolysis technologies. Energies. 2018;11:3115
  41. 41. Luque R, Menendez JA, Arenillas A, Cot J. Microwave-assisted pyrolysis of biomass feedstocks: The way forward? Energy & Environmental Science. 2012;5:5481-5488
  42. 42. Yuan T, He W, Yin G, Xu S. Comparison of bio-chars formation derived from fast and slow pyrolysis of walnut shell. Fuel. 2020;261:116450
  43. 43. Brown TR, Wright MM, Brown RC. Estimating profitability of two biochar production scenarios: Slow pyrolysis vs fast pyrolysis. Biofuels, Bioproducts and Biorefining. 2011;5(1):54-68
  44. 44. Daful AG, Chandraratne MR. Biochar production from biomass waste-derived material. In: Choudhury I, Hashmi S, editors. Encyclopedia of Renewable and Sustainable Materials. 1st ed. New York: Elsevier; 2018. pp. 370-378
  45. 45. Oochit D, Selvarajoo A, Arumugasamy SK. Pyrolysis of biomass. In: Singh L, Kalia VC, editors. Waste Biomass Management—A Holistic Approach. New York: Springer; 2017. pp. 215-229
  46. 46. Kan T, Strezov V, Evans TJ. Lignocellulosic biomass pyrolysis: A review of product properties and effects of pyrolysis parameters. Renewable and Sustainable Energy Reviews. 2016;57:1126-1140
  47. 47. Foong SY, Liew RK, Yang Y, Cheng YW, Yek PNY, Mahari WAW, et al. Valorization of biomass waste to engineered activated biochar by microwave pyrolysis: Progress, challenges, and future directions. Chemical Engineering Journal. 2020;389:124401
  48. 48. Tripathi M, Sahu JN, Ganesan P. Effect of process parameters on production of biochar from biomass waste through pyrolysis: A review. Renewable and Sustainable Energy Reviews. 2016;55:467-481
  49. 49. Chen Z, Wang M, Ren Y, Jiang E, Jiang Y, Li W. Biomass torrefaction: A promising pretreatment technology for biomass utilization. IOP Conference Series: Earth and Environmental Science. 2018;113:012201
  50. 50. Bamdad H, Hawboldt K, MacQuarrie S. A review on common adsorbents for acid gases removal: Focus on biochar. Renewable and Sustainable Energy Reviews. 2018;81:1705-1720
  51. 51. Zhu X, Luo Z, Diao R, Zhu X. Combining torrefaction pretreatment and co-pyrolysis to upgrade biochar derived from bio-oil distillation residue and walnut shell. Energy Conversion and Management. 2019;199:111970
  52. 52. Chen D, Mei J, Li H, Li Y, Lu M, Ma T, et al. Combined pretreatment with torrefaction and washing using torrefaction liquid products to yield upgraded biomass and pyrolysis products. Bioresource Technology. 2017;228:62-68
  53. 53. Zeng K, He X, Yang H, Wang X, Chen H. The effect of combined pretreatments on the pyrolysis of corn stalk. Bioresource Technology. 2019;281:309-317
  54. 54. Aliyu AS, Abdullahi N, Sulaiman F. Pyrolysis of torrefied oil palm wastes for better biochar. Malaysian Journal of Fundamental and Applied Sciences. 2017;13:124-128
  55. 55. Zhang Z, Zhu Z, Shen B, Liu L. Insights into biochar and hydrochar production and applications: A review. Energy. 2019;171:581-598
  56. 56. Lee J, Sarmah AK, Kwon EE. Production and formation of biochar. In: Ok YS, Tsang DC, Bolan N, Novak JM, editors. Biochar from Biomass and Waste: Fundamentals and Applications. Amsterdam: Elsevier; 2019. pp. 3-18
  57. 57. Gan YY, Ong HC, Show PL, Ling TC, Chen W-H, Yu KL, et al. Torrefaction of microalgal biochar as potential coal fuel and application as bio-adsorbent. Energy Conversion and Management. 2018;165:152-162
  58. 58. Saqib NU, Sharma HB, Baroutian S, Dubey B, Sarmah AK. Valorisation of food waste via hydrothermal carbonisation and techno-economic feasibility assessment. The Science of the Total Environment. 2019;690:261-276
  59. 59. Oktaviananda C, Rahmawati RF, Prasetya A, Purnomo CW, Yuliansyah AT, Cahyono RB. Effect of temperature and biomass-water ratio to yield and product characteristics of hydrothermal treatment of biomass. AIP Conference Proceedings. 1823;2017:1-7
  60. 60. Polprasert C, Koottatep T. Bioenergy production. In: Polprasert C, Koottatep T, editors. Organic Waste Recycling. Technology, Management and Sustainability. London: IWA Publishing; 2017. pp. 157-251
  61. 61. Krylova AY, Zaitchenko V. Hydrothermal carbonization of biomass: A review. Solid Fuel Chemistry. 2018;52:91-103
  62. 62. Guiotoku M, Maia CM, Rambo CR, Hotza D. Synthesis of carbon-based materials by microwave hydrothermal processing. In: Chandra U, editor. Microwave Heating. New York: IntechOpen; 2011. pp. 291-308
  63. 63. Xu Y, Xia M, Jiang Y, Li F, Xue B. Opal promotes hydrothermal carbonization of hydroxypropyl methyl cellulose and formation of carbon nanospheres. RSC Advances. 2018;8:20095-20107
  64. 64. Libra JA, Ro KS, Kammann C, Funke A, Berge ND, Neubauer Y, et al. Hydrothermal carbonization of biomass residuals: A comparative review of the chemistry, processes and applications of wet and dry pyrolysis. Biofuels. 2011;2:71-106
  65. 65. Wiedner K, Rumpel C, Steiner C, Pozzi A, Maas R, Glaser B. Chemical evaluation of chars produced by thermochemical conversion (gasification, pyrolysis and hydrothermal carbonization) of agro-industrial biomass on a commercial scale. Biomass and Bioenergy. 2013;59:264-278
  66. 66. Kwapinski W. Char production technology. In: Jeguirim M, Limousy L, editors. Char and Carbon Materials Derived from Biomass: Production, Characterization and Applications. Amsterdam: Elsevier; 2019. pp. 39-68
  67. 67. Wang X, Chi Q, Liu X, Wang Y. Influence of pyrolysis temperature on characteristics and environmental risk of heavy metals in pyrolyzed biochar made from hydrothermally treated sewage sludge. Chemosphere. 2019;216:698-706
  68. 68. Olszewski M, Arauzo P, Wądrzyk M, Kruse A. Py-GC-MS of hydrochars produced from brewer’s spent grains. Journal of Analytical and Applied Pyrolysis. 2019;140:255-263
  69. 69. Garlapalli RK, Wirth B, Reza MT. Pyrolysis of hydrochar from digestate: Effect of hydrothermal carbonization and pyrolysis temperatures on pyrochar formation. Bioresource Technology. 2016;220:168-174
  70. 70. Yang X, Tsibart A, Nam H, Hur J, El-Naggar A, Tack FM, et al. Effect of gasification biochar application on soil quality: Trace metal behavior, microbial community, and soil dissolved organic matter. Journal of Hazardous Materials. 2019;365:684-694
  71. 71. Wang J, Wang S. Preparation, modification and environmental application of biochar: A review. Journal of Cleaner Production. 2019;227:1002-1022
  72. 72. Igalavithana AD, Mandal S, Niazi NK, Vithanage M, Parikh SJ, Mukome FN, et al. Advances and future directions of biochar characterization methods and applications. Critical Reviews in Environmental Science and Technology. 2017;47:2275-2330
  73. 73. Wang T, Zhai Y, Zhu Y, Li C, Zeng G. A review of the hydrothermal carbonization of biomass waste for hydrochar formation: Process conditions, fundamentals, and physicochemical properties. Renewable and Sustainable Energy Reviews. 2018;90:223-247
  74. 74. You S, Ok YS, Chen SS, Tsang DC, Kwon EE, Lee J, et al. A critical review on sustainable biochar system through gasification: Energy and environmental applications. Bioresource Technology. 2017;246:242-253
  75. 75. Fryda L, Visser R. Biochar for soil improvement: Evaluation of biochar from gasification and slow pyrolysis. Agriculture. 2015;5:1076-1115
  76. 76. Ndirangu SM, Liu Y, Xu K, Song S. Risk evaluation of pyrolyzed biochar from multiple wastes. Journal of Chemistry. 2019;2019:1-28
  77. 77. Meyer S, Genesio L, Vogel I, Schmidt H-P, Soja G, Someus E, et al. Biochar standardization and legislation harmonization. Journal of Environmental Engineering and Landscape Management. 2017;25:175-191
  78. 78. Llorach-Massana P, Lopez-Capel E, Peña J, Rieradevall J, Montero JI, Puy N. Technical feasibility and carbon footprint of biochar co-production with tomato plant residue. Waste Management. 2017;67:121-130
  79. 79. Cowie AL, Downie AE, George BH, Singh B-P, Van Zwieten L, O'Connell D. Is sustainability certification for biochar the answer to environmental risks? Pesquisa Agropecuária Brasileira. 2012;47:637-648
  80. 80. Gelardi DL, Li C, Parikh SJ. An emerging environmental concern: Biochar-induced dust emissions and their potentially properties. The Science of the Total Environment. 2019;678:813-820

Written By

Nor Adilla Rashidi and Suzana Yusup

Submitted: 24 December 2019 Reviewed: 22 April 2020 Published: 22 May 2020