Open access peer-reviewed chapter

Biochar: Production, Application and the Future

Written By

Edward Kwaku Armah, Maggie Chetty, Jeremiah Adebisi Adedeji, Denzil Erwin Estrice, Boldwin Mutsvene, Nikita Singh and Zikhona Tshemese

Submitted: 19 April 2022 Reviewed: 26 April 2022 Published: 06 June 2022

DOI: 10.5772/intechopen.105070

From the Edited Volume

Biochar - Productive Technologies, Properties and Applications

Edited by Mattia Bartoli, Mauro Giorcelli and Alberto Tagliaferro

Chapter metrics overview

1,308 Chapter Downloads

View Full Metrics

Abstract

Biochar, or carbon obtained from biomass, is a particularly rich source of carbon created by thermal burning of biomass. There is a rise of interest in using biochar made from waste biomass in a variety of disciplines to address the most pressing environmental challenges. This chapter will provide an overview on the methods employed for the production of biochar. Biochar has been considered by a number of analysts as a means of improving their ability to remediate pollutants. Process factors with regards to biochar properties are mostly responsible for determining biomass production which is discussed in this present chapter. Several characterization techniques which have been employed in previous studies have received increasing recognition. These includes the use of the Fourier transform infrared spectroscopy and the Scanning electron microscope which duly presented in this chapter. This chapter also discusses the knowledge gaps and future perspectives in adopting biochar to remediate harmful contaminants, which can inform governmental bodies and law-makers to make informed decisions on adopting this residue.

Keywords

  • biochar
  • biomass
  • characterization
  • future perspective
  • pyrolysis
  • pretreatment

1. Introduction

The word char, is a common terminology used for the solid product of the combustion of carbonaceous material [1]. Generally, char product is rich in carbon content; an example is charcoal, which is almost the earliest invention of humans from fire or heat creation. Another vivid example of char is biochar. In this case, the study, is made from organic compounds such as forest, agricultural or animal products but in the absence/limited supply of oxygen compared to charcoal. Therefore, biochar is derived from biomass combustion in the presence of a limited oxygen supply and at relatively low temperatures below 700°C. The earliest known purpose for creating biochar was specifically for soil application such as carbon storage or sequestration in soil; improvement of soil performance such as increase in nutrient availability, reduction of compactness in soil, soil pH improvement; soil water filtration. Recent applications involve energy production, biochemical process stability and improvement, climate change mitigation, and construction additive [1, 2, 3]. The raw material determines carbonized organic matter properties and the operational parameters used during it production. Pyrolysis (slow or fast) and gasification are the main methods for the production of biochar. The physical nature of the biochar produced is directly affected by the chemical composition of the biomass feedstock. Most organic matter begin to thermally decompose at temperatures above 120°C. Hemicelluloses degrade between 200 and 260°C, cellulose between 240 and 350°C, and lignin between 280 and 500°C. As a result, the proportions of these components will affect the degree of reactivity and, as a result, the extent to which the physical structure is modified during processing [4]. Biochar is characterized with high porosity with pores ranging in size from micro to macropores. Large holes, which originate from the raw biomass’s vascular bundles, are critical for increasing soil quality because they can serve as habitats for symbiotic microbes. Biochar major components are carbon, volatile matter, mineral matter (ash), and moisture. The percentage composition of each components varies based on the feedstock material and the operating parameters [1]. Biochar from plant-based materials have higher carbon composition which range from as low as 51% to as high. The understanding of the key mechanisms for changes in physicochemical properties of biochar during processing for various feedstock types and operating parameters is required to determine biochar’s potential for application both now and in future. Therefore, this chapter explains biochar production techniques, factors affecting its properties and compositions and its application.

Advertisement

2. Biochar production techniques

An ever-growing appetency for using biochar for various applications has orchestrated an increase in converting it into biochar. Thermochemical conversion is a common technology for making biochar. Thermochemical conversion techniques are pyrolysis, hydrothermal carbonization (HTC), gasification, torrefaction, and hydrothermal liquefaction [5, 6].

2.1 Pyrolysis

Pyrolysis is a thermochemical technique that produces biochar, bio-oil, and syngas derived from biomass [7]. The process involves heating and thermally decomposing biomass under anaerobic conditions or limited oxygen supply (low stoichiometric oxygen atmosphere) with temperatures ranging between 400°C and 1200°C [2]. The absence of oxygen enables biomass heating beyond its thermal stability limit, causing the creation of more robust products, including solid residues. By creating an anaerobic atmosphere, it is also ensured that combustion will not occur when the biomass is heated. It is a highly complex process involving many distinct reactions in the reacting zone [8]. In another study, a low-temperature range for pyrolysis was recorded between 250°C and 900°C. Biomass from Agriculture comprises lignin, cellulose, hemicelluloses, and silica. Typically, cellulose pyrolyzes at 350°C, whereas the melting point of lignin is well above 350°C [6]. Although the product yield depends on various operating variables, char formation is generally favored by low temperatures and long residence times [9]. Therefore, it can be decoded that the effective temperature range for pyrolysis was between 300 and 700°C. The cracking of heavy chemicals happens in secondary pyrolysis and converts biomass into biochar or gases. Figure 1 is a summary of the pyrolysis technique and the operating variables affecting pyrolysis.

Figure 1.

Schematic representation of pyrolysis process [3].

In essence, this is an alternative way to valorize biomass into various products such as bio-oil, syngas and biochar. Depolymerization, fragmentation, and cross-linking are chemical mechanisms that occur during the process at specific temperature points, resulting in a different product state for lignocellulosic components, including cellulose and hemicellulose (solid, liquid and gas). Biochar and bio-oil are the solid and liquid products, whereas CO2, CO, H2, (collectively known as syngas) are evolved as the gaseous by-products (C1-C2 hydrocarbons) [3]. Biochar is made in a different type of reactors, such as paddle kiln, bubbling fluidized bed, wagon reactor, and agitated sand rotating kiln. The biomass nature and employed type determine the biochar yield during the pyrolysis route. The major operating parameter that impacts product efficiency is the temperature [10, 11]. When the pyrolysis temperature is increased, biochar’s yield decreases and the generation of syngas increases. The gas yield is represented by the initial section of the product side (as shown in Eq. (1)), with various gases created during the process.

C6H6O6nH2+CO+CH4++C5H12+H2O+CH3OH+CH3COOH++CE1

The mixture of multiple sorts of liquid outputs is shown in the second part of the products’ side, and the solid yield is represented in the last component [12]. One of the most significant masteries of this technology is that it may be optimized to achieve the desired outcomes. Slow pyrolysis, for example, can be utilized to produce a considerable amount of biochar, whereas fast pyrolysis is better for dominantly producing bio-oil [13].

2.1.1 Types of pyrolysis

Pyrolysis is strongly dependent on the operating parameters, namely temperature, heating rate, and residence time [14]. These operating conditions further help to categorize pyrolysis into other six subclasses. These subclasses are slow pyrolysis, fast pyrolysis, flash pyrolysis, vacuum pyrolysis, intermediate pyrolysis, and hydropyrolysis [15]. Each classification of pyrolysis has its own documented benefits and drawbacks. The subclasses in question foster an environment for different reaction conditions and mechanisms to have various products. The pyrolysis technology mechanism is shown in Figure 2.

Figure 2.

Representation of a pyrolysis process [6].

2.1.1.1 Slow pyrolysis

As indicated by the name, to complete the process, slow pyrolysis has a long residence time (more than 1 hour), and biochar is produced as a major product [16]. Slow pyrolysis is dubbed conventional pyrolysis, where biomass is heated at temperatures ranging between 300 and 600°C accompanied by a heating rate of 5–7°C/min [12, 17]. A lower heating rate and longer vapor residence time provide a suitable environment and adequate time for the secondary reactions to proceed. Furthermore, a prolonged residence period permits vapors created during the secondary reaction to be evacuated [15, 18]. This leads to the creation of solid carbonaceous biochar in the end. Slow pyrolysis favors char development, but liquid and gaseous products are also created in modest quantities. Biochar is formed as a primary product (35–45%) together with other products such as bio-oil (25–35%) and syngas (20–30%), as indicated in Eq. (1) [6, 19].

2.1.1.2 Fast pyrolysis

Fast pyrolysis is a direct thermochemical process for converting solid biomass into high-energy liquid bio-oil. A high-efficiency thermochemical technique to produce biomass-derived biofuels, with reduced amounts of solids and gases produced [20, 21]. Fast pyrolysis is carried out without oxygen at temperatures above 500°C and a heating rate of over 300°C/min. Fast pyrolysis is a rapid biochar generation technique that takes only a few seconds. Fast pyrolysis produces 60% bio-oil, 20% biochar, and 20% syngas, as reported in other studies [21, 22]. Even higher temperatures in the range of 850–1250°C with a heating rate of 10–200°C for a short residence time ranging from 1 to 10 s have been reported in several experiments. 60%-75% of liquid products, 15%-25% of biochar and 10–20% of non-condensable gaseous products are produced by a typical pyrolysis process [23]. Fast pyrolysis takes biomass to temperatures in which thermal cracking can occur and minimizes the exposure time, which supports biochar production [24].

2.1.1.3 Flash pyrolysis

This is dubbed to be an enhanced and modified version of fast pyrolysis. Biomass decomposes quickly, usually in less than a minute, at 1000°C and even higher temperatures. Heating rates of above 1000°C/sec have been recorded on occasion. Flash pyrolysis is carried out at temperatures between 900 and 1200°C, which can be reached in less than one second (usually between 0.1 and 1 s) [25, 26]. A high bio-oil yield combines a high heating rate with a high temperature and a short vapor residence time. However, the yield of biochar is reduced because of the process [27, 28]. In flash pyrolysis, heat and mass transfer processes, reaction chemical kinetics and biomass phase transition behavior all play a role in product distribution. Although flash pyrolysis is performed in a fluidised bed reactor and a twin-screw mixing reactor, it has limited industrial applicability because of the reactor’s architecture, which requires it to run at a high temperature with a very high heating rate [12].

2.1.1.4 Vacuum pyrolysis

This is the thermal decomposition of biomass under vacuum or relatively low pressure in an isolated oxygen environment [15, 29]. Pressure is usually regulated in the region between 0.5 and 2 bar, and temperature is maintained at 450–600°C [30]. Like slow pyrolysis, vacuum pyrolysis has comparably low heating rates. However, these two techniques, in comparison, yield significantly different products. This owes to the constant and effective discharge of the vapor produced during vacuum pyrolysis through condensation train. The rapid evacuation of organic vapors created during the primary pyrolysis also considerably minimizes the vapor residence time, which in turn minimizes the occurrence of secondary reactions and assures a high liquid product yield during the secondary pyrolysis [31]. As a result, only vacuum or low-pressure extraction is utilized to remove vapor evolved during pyrolysis, which substantially affects product quality and yield by preventing inorganic devolatilisation.

2.1.1.5 Intermediate pyrolysis

As the name suggests, this is a combination of slow and fast pyrolysis processes, and it is crucial when there is a need to balance solid and liquid products. This means that slow pyrolysis is more efficient at producing large amounts of char, but it also results in lower amounts of liquid products, while it is vice versa with fast pyrolysis. Generally, pressure is kept at 1 bar during the process. Intermediate pyrolysis has temperatures ranging between 500 and 650°C, with heating rates between 0.1 and 10°C/min and residence time between 5 and 17 mins [32]. 40–60% liquid, 20–30% non-condensable gases, and 15–25% biochar are typical constituents of finished products [33, 34]. Using intermediate pyrolysis conditions prevents the synthesis of high molecular reactive tars and results in dry biochar, which can be utilized for agricultural purposes or directly in boilers and engines in conjunction with high-quality bio-oil [2].

2.1.1.6 Hydropyrolysis

It relatively a novel technique that is used for the conversion of biomass into high quality products by injection of hydrogen or hydrogen based material into the reactor under high pressure, typically above the atmospheric pressure, stretching from 50 bar to 200 bar [15, 35]. The heating rate (10–300°C/s), residence time (over 15 sec) and temperature (350–600°C) are not highly deviated from fast pyrolysis [36]. In essence, hydropyrolysis can be considered a special type of fast pyrolysis subjected to high pressure in an atmosphere infused with hydrogen or hydrogen-based material. This method is not ideal for the production of biochar as the introduction of hydrogen under high temperature and pressure acts as a reducing agent, hence reducing oxygen content in the bio-oils produced and synchronously inhibiting the production of biochar [37, 38]. The employment of a catalyst to eradicate oxygen, water, and COx from the liquid product is typically linked with hydropyrolysis. Catalysts also reduce depolymerisation and coking reactions [39]. However, developing the catalyst for this intention remains a notable example of the difficult aspects of catalytic hydropyrolysis.

2.2 Carbohydrate decomposition

The majority of the material used in biochar production via pyrolysis contain carbohydrates in various forms (cellulose, hemicellulose and lignin), and these react differently based on the operating conditions they are subjected to, thus influencing the product yield of pyrolysis [15]. More specifically, lignin and cellulose are the major parts of biomass, making up its bulk [40]. On pyrolysis, cellulose mostly creates tar, a mixture of discrete ketones, aldehydes, organic liquids, and char, whereas lignin essentially produces char and a minimal amount of water. As the cellulose content grows but the char and tar content decreases, the yield of gaseous content increases. It’s also been discovered that structural differences in biomass cause changes in the pyrolysis product’s composition [41].

2.2.1 Cellulose decomposition

By lowering the extent of polymerization, the process of cellulose degradation is determined, which consists of two principal reactions:

  1. Slow pyrolysis involves cellulose degradation over a prolonged period with a lower heating rate.

  2. Fast pyrolysis occurs at high heating rates through speedy volatilization and leads to levoglucosan formation.

In addition to producing the solid product biochar, levoglucosan is dehydrated to generate hydroxymethylfurfural, which can break down to produce liquid and gaseous products such as bio-oil and syngas, respectively. Furthermore, the hydroxymethylfurfural can undergo several processes, including aromatization, condensation, and polymerization, to generate solid biochar [42, 43]. At low temperatures, cellulose degrades to a reasonably stable anhydrocellulose that produces a lot of char, but it decomposes into volatiles [25, 44].

2.2.2 Hemicellulose decomposition

The hemicellulose degradation mechanism is like that of cellulose. Depolymerisation of hemicellulose leads to oligosaccharides production [45]. Decarboxylation, intramolecular rearrangement, depolymerisation, and aromatisation reactions can be used to synthesize biochar or the compound can degrade into syngas and bio-oil [46]. The volatile products and lignin are responsible for the char yield of the cellulose and hemicellulose components in biomass [40].

2.2.3 Lignin decomposition

Unlike the degradation of cellulose and hemicellulose, lignin decomposition is more complicated [47]. The creation of a more condensed solid structure and the shattering of relatively weak bonds result in the formation of char from lignin [48]. The β-O-4 lignin bond is broken and causes free radicals to be released. The protons emanating from other particles are captured by these free radicals, causing the production of degraded substances or compounds. Chain propagation is accomplished by free radicals moving to other molecules. Different amounts of lignin related to variable wood types bring about different breakdown rates. Coniferous lignin has been discovered to be more stable than deciduous lignin, and the former creates more char [49, 50].

2.3 Gasification

This is a thermochemical process that decomposes carbon-rich materials into gaseous products, including CO, CO2, CH4, H2, and traces of hydrocarbons; these gases are referred to as syngas [51, 52]. Gasification happens at high temperatures between 700 and 900°C in an environment with restricted oxidizing agents such as oxygen, air, nitrogen, steam, carbon dioxide, or a mixture of these gases. It was discovered that when the temperature rose, carbon monoxide and hydrogen production increased, while other components such as methane, carbon dioxide, and hydrocarbons declined [53]. The main product of this process is syngas (mostly hydrogen), while char is referred to as a by-product (or waste) with a lower yield, along with ash, tar, and some oil [51]. Partial oxidation of biomass, unlike combustion, takes the energy available in the biomass and bundles it into chemical bonds in the form of gaseous products. The intrinsic chemical energy of carbon in biomass is transformed into combustible fuel gases, which are more efficient and convenient to utilize than raw biomass [54]. Commercial use of the gasification technique has also been documented. Because of its lower Levelised emissions and higher volume of syngas, gasification outperforms other traditional techniques including pyrolysis, combustion, and fermentation. The O/C ratio is critical to achieving high gasification efficiency. High gasification efficiency is achieved by using biomass with a low O/C ratio during gasification. Biomass can be reduced in its O/C ratio by the process of torrefaction. Before conventional gasification, torrefaction might be regarded as a pretreatment for better product quality. It is a low-temperature process between 200 and 300°C with a heating rate of roughly 50°C/min depending on the biomass composition and type [55, 56]. Pyrolysis and gasification are closely related processes. When gasification and pyrolysis are combined, there is no apparent separation between the two approaches [57, 58]. The little composition of oxygen used in gasification causes the biomass to undergo partial oxidation, changing the final product’s characteristics. The product type is one of the most significant variations between pyrolysis and gasification. Gasification produces around 85% gaseous products, 10% solid char, and 5% liquid products [15, 58]. The schematic of the gasification process is shown in Figure 3.

Figure 3.

Process diagram for gasification [54].

The gasification mechanism can be sub-divided into many steps as follows [5]:

2.3.1 Drying

Biomass moisture is entirely removed from the material, and no energy is recovered in the process. Different types of biomass have varying moisture contents. When the biomass has a high moisture content, drying is used as a distinct step during gasification.

2.3.2 Pyrolysis

The biomass is heated from 200 to 700°C with restricted oxygen or air during the pyrolysis process. The volatile components of the biomass are evaporated under these circumstances. The volatile vapor contains CO, CO2, CH4, H2, tar (heavier hydrocarbon) gases, and water vapor [59]. Tar and char are also formed [60].

2.3.3 Oxidation/combustion

The oxidation and combustion reactions of the gasification agents are the primary energy sources for the gasification process. These gasification agents react with the gasifier’s combustible species to create CO2, CO, and water.

2.3.4 Reduction

The CO2 and H2O are produced when the oxygen provided to the gasifier combines with the combustible elements. Upon contact with the char formed by pyrolysis, some of this CO2 and H2O are converted to CO and H2 [60, 61]. Furthermore, the hydrogen in the biomass can be oxidized, resulting in the production of water. The reduction reactions that take place inside the gasifier are endothermic, and the energy necessary for them comes from the combustion of char and volatiles. Through a series of reactions, biomass reduction produces combustible gases such as hydrogen, carbon monoxide, and methane [62, 63].

2.3.5 Cracking

Furthermore, during the gasification process, the tar gases formed during the pyrolysis step are cracked, resulting in non-condensable gasses, light hydrocarbons, and unconverted tar [64]. The cracking stage follows more or less Eq. (2).

aCnHxbCmHy+CH2E2

Where CnHx is tar and CmHy is dehydrogenated hydrocarbons; a, b and c are mole ratios.

Advertisement

3. Factors affecting the properties of biochar

3.1 Feedstock

Biomass is a composite solid substance made up of organic, inorganic and biological material produced from living or non-living creatures/organisms. There are two main categories of biomass, namely Woody and Non-woody biomass. Woody biomass is mainly forestry and tree residue [1]. It is characterized by low moisture and ash content, high calorific and bulk density values, and low voidage; in contrast, Non-woody biomass is made up of agricultural crop residue, animal waste, and municipal and industrial solid waste [1]. Non-woody biomass is characterized by high moisture and ash content, decreased calorific value, low bulk density, and increased voidage compared to woody biomass [1]. The moisture content of the biomass has been shown to have a significant effect on the physicochemical characteristics of the derived biochar [2]. A study conducted by [3] comparing the pyrolytic charcoals produced from hard and softwood bark samples reported a direct correlation between initial sample moisture content and the surface chemistry derived charcoal; the study found that a decrease in the moisture content of maple bark resulted in charcoal surface becoming more graphite-like and polyaromatic attributed to prolonged pyrolysis time. The effect of feedstock lignin and cellulose content on biochar formation is a well-researched area [4]. Lignin is an amorphous, high molecular weight polymer that is hydrophobic in nature and has several aromatic functional groups in comparison; cellulose and hemicelluloses are made up of simple sugar monomers that disintegrate at temperatures below 450 degrees Celsius [5]. Studies conducted by Tripathi et al. 2016 and Yu et al. 2014 [2, 6] showed that the cellulose content of feedstock aided the formation of tar (which comprises aldehydes, organic liquids, ketones, and char); while a high lignin concentration is beneficial to the formation of char during pyrolysis. According to Demirba (2004) [7], high feedstocks lignin content will increase char formation. It has been shown that increased lignin content in plant biomass promotes carbonization and increases biochar carbon and ash content [8, 9].

3.2 Residence time

Residence (pyrolysis time) has been shown to affect the degree of carbonization and biochar yield of feedstock; this effect is particularly pronounced at low temperatures [18]. According to Zornoza et al. (2016), increased residence time during pyrolysis results in a higher degree of carbonization, reducing the liable organic matter mitigation the vulnerability of the biochar to microbial attack [19]. Residence time has also been shown to influence the specific surface area of biochar produced. A study conducted by Wang et al. (2019) found that the surface area of biochar’s derived from the co-pyrolysis of sewage sludge and cotton stalks increased as residence time increased from 30 minutes to 90 minutes [20]. This was attributed to the formation and extension of pore structures of the biochar caused by the increased thermal decomposition of organic matter and volatiles released from etching pores during the increased residence time [21]. The same study noted a decrease in the surface area of the biochar’s as the residence time was increased from 90 minutes to 150 minutes; this reduction was accounted for by the collapse of the pore structure of the biochar during the extended residence time [20]. Residence time has also been shown to affect the calorific value of the biochar produced; a study conducted by Ahmad et al. (2020) on coconut shell derived biochar showed an increase in calorific value from 25.99 MJ/kg to 29.54 MJ/kg as residence time increased for 45 minutes to 75 minutes [22].

3.3 Biomass pretreatment

The pre-treatment of biomass before the pyrolysis has been shown to influence biochar characteristics. Pre-treatment is primarily divided into four categories: physical, physiochemical/thermal, chemical, and biological. Physical pre-treatment describes methods (milling, grinding etc.) that use mechanical energy to alter biomass properties. The most common form of physical pre-treatment is particle size reduction via mechanical comminutions. The effect of particle size reduction and fractionation of ash content is well researched. A study conducted by Liu et al. showed that the ash content of switchgrass and pine bark varied considerably with particle size fractions [22]. The study also reported the potential 20% removal of inorganic constituents from switchgrass and a 30% removal of inorganic constituents from raw pine bark. A similar study conducted by Bridgeman et al. found that the ash content of switchgrass and reed canary greatly increased in fines with particle sizes smaller than 90 micrometers, increasing to 3.62 wt. % to 6.0 wt. % for reed canary grass and 3.12 wt. % to 6.88 wt. % (dry basis) for switchgrass [23]. Besides the ash content, feedstock particle size is also correlated to biochar particle size, with finer feedstocks producing finer biochar particle sizes [18]. Studies have found that biochar’s derived from finer feedstocks exhibit lower nitrogen content as well as increased surface area, electrical conductivity, and pH [24, 25]. A study conducted by Sun et al. (2012) evaluating the properties of fine apple wood and corn stover-derived biochar (feedstock = 0.25 mm) reported a higher surface area when compared to applewood or corn stover-derived biochar stover-derived biochar of feedstock particle size = 1.5 mm [27]. Thermal pre-treatment describes methods that make use of thermal energy to produce changes in biomass properties; the most common forms of thermal pre-treatment are steam explosion, HTC and hot water extraction. Steam explosion involves the subjection of biomass to high temperatures and pressures between (160-260°C) and (0.69–4.83 MPa); the biomass subsequently undergoes sudden decompression scattering the fiber material and breaking the covalent bonds between the hemicellulose and lignin [28, 29]. Steam explosion increases the lignin content of the biomass by facilitating the depolymerisation of lignin into lower molecular weight molecules, which then condense with other degradation products [30]. A study conducted by Chen et al. 2017 [46] evaluating the effect of the steam explosion of crop straws before pyrolysis reported a change in the surface structure of the derived biochar; exhibiting a rougher surface when compared to the smoother, clearer and distinct pore structure of the untreated crop straw [31]. The same study also showed an approximate increase in the specific surface area of oil-rape straw-derived biochar 16 times greater than that on the untreated sample.

Advertisement

4. Biochar characterization

Properties of biochar produced depend on the composition, type of biomass and the conditions under which it is carbonized. Both physical and chemical characterizations are necessary when identifying the basic properties of biochar and predicting the various application uses. Biochar serves as a promising alternative to its surface area, charged surfaces and functional groups. Figure 4 below displays the different physical and chemical methods used for biochar characterization, focusing on BET and FTIR, belonging to the chemical characterization and SEM as physical characterization.

Figure 4.

Overview of a proposed characterization techniques for biochar [65].

The main aim of quantification to distinguish biochar from organic matter and other forms of black carbon produced. Majority of the potential technology is dependent on spectroscopic characteristics rather than physical separation or isolation.

Biochar being produced from a range of biomass that has different chemical and physical properties results in materials of different properties. Properties of each biomass are important during thermal conversion processes, proximate analysis (ash and moisture content); calorific value; fractions of fixed carbon; volatile components; fractions of lignin, cellulose and hemicellulose; inorganic substances; true density; particle size and moisture content.

4.1 Porosity and surface area

Chemical composition of biomass feedstock and biomass is subjected to a range of analyses to achieve the basic physicochemical characteristics of each raw material. Figure 5displays the physiochemical characteristics of biochar. Biochar production is often assessed through changes in the elemental concentrations of C, H, O, S and N and the associated ratios. The fixed carbon is the solid residue that remains after the particle size is carbonized and the volatile matter is expelled. The H/C and O/C ratios are used to determine the degree of aromaticity and maturation. Elemental ratios of O/C, O/H and C/H have been used to provide a reliable measure of the extent of pyrolysis and the level of oxidative adjustment of the biochar. Irrespective of the pyrolytic temperature, the BET areas increased with an increase in carbon burn off, indicating that the carbon burns off had a significant role in increasing pore volume and surface area while the average pore size increased with residence time and pyrolytic temperature. The BET surface area of biochar value of (1057 m2 .g − 1) has been reported, which appears slightly higher than that of activated carbon (970m2 .g−1). Biochar micropore volume of (0.24 mL .g−1) also appeared smaller than that of activated carbon, having a value of (0.32 mL .g−1), however having an average pore diameter of (5.2 nm).

Figure 5.

Fourier-transform infrared spectra (FTIR) of the biochar samples [66].

4.2 Scanning electron microscope (SEM)

Scanning electron microscopy is categorized as a physical characterization technique used to determine the samples macroporosity and the physical morphology of solid substance (Figure 6). A study by Amin 2016 [1] approximated that the biochar produced from cellulose plant materials had a pore diameter of 1 𝜇m. This characteristic is highly dependable in the intrinsic architecture of the feedstock use.

Figure 6.

SEM micrograph of biochar with magnification of 500x [67].

SEM micrographs displayed that the biochar produced at different pyrolytic temperatures has a distinguishable and clear honeycomb structural appearance due to the original tubular structures present in plant cell materials (Figure 6). The well-developed pores have a direct impact on the high surface area. According to Cantrell et al. (2012), biochar produced at lower temperatures is appropriate for regulating fertilizer nutrients and absorbing pollutants from the soil. Higher temperatures lead to material analogous to activated carbon and environmental remediation. SEM micrographs of biochar displayed a clean surface as the pyrolysis process had stabilized the volatile hydrocarbons, therefore smoothening the surface of the biochar. Pyrolysis at lower temperatures displays molded structures with small pores and uneven surface structure. In general, it is safe to say that since the biomass wastes contain lignin and high volatile matter content, the pore creation in biochar is directly affected.

4.3 Fourier transform infrared spectroscopy (FTIR)

FTIR spectroscopy serves as a great tool to observe the shift change of chemical compositions. The commonly used technique for biochar characterization using the FTIR is the pellet technique, which mixes 1 mg of dried biochar with 300 mg of pre-dried and pulverized spectroscopic grade KBr. Novak ae al. (2012) used the pellet technique to conclude 3400to 3410 cm−1, H-bonded O–H stretching vibrations of hydroxyl groups from alcohols, phenols, and organic acids, 2850 to 2950 cm−1, C–H stretching of alkyl structures; 1620–1650 cm−1, aromatic and olefinic CDC vibrations, CDO in amide (I), ketone, and quinone groups; 1580 to 1590 cm−1, COO- asymmetric stretching; 1460 cm−1, C–H deformation of CH3 group; 1280–1270 cm−1, O–H stretching of phenolic compounds; and three bands around 460, 800, and 1000–1100 cm−1, bending of Si–O stretching [68]. Figure 5 illustrates the FTIR spectra of biochar collected during different stages of the production, i.e. (Biochar: Original, −1: pre-incubation, −2: jointing, −3: Heading; −4: Mature).

Advertisement

5. Applications of biochar and future perspective

Biochar is a product (together with bio-oil and gases) resulting from biomass pyrolysis. Biochar usage has increased because it reduces the negative impacts of biomass on the environment [69]. The physicochemical properties of biochar are what govern the applications of this material. Depending on the feedstock type, production technology and process conditions [70]; the quality, yield and toxicity of the resulting biochar differs (as shown in Table 1) [72, 73]. These applications (including potential applications) range from adsorption for water and air pollutants [74], activated carbon [75], anaerobic digestion promoter/catalyst [76], construction material [77], agriculture and horticulture use such as soil conditioning, compost additive [78], carbon sequestration, etc. [73]. Figure 7 demonstrates these applications and how biochar contributes to the circular economy through its uses in agriculture and horticulture. Also, these numerous biochar benefits show a great potential to contribute to the economic sustainability of emerging cellulosic bioenergy production systems [79, 80]. It is worth noting that as the number of applications of biochar increases, so does the number of manufacturers, leading to a need for regulated standards and guidelines for the production of this material (see Table 2) [81, 82].

Type of characterizationDetermination methodResults and remarks
Elemental analysisC, H, O, S and N associated ratiosThe H/C, O/C and N ratios are used to determine the aromaticity and maturity of the biochar
BETSurface area, pore structure, average pore diameter, pore volume and average pores of biochar1057m2g−1; macroporosity and microporosity; 5.2 nm; 0.24 mLg−1; 3.3 nm. [71]
FTIRChanges which occur in the biochar preparations as well as its functional groups present from the original biochar.Changes include dehydration, pyrolysis, graphene
nucleation, and finally carbonization; O–H
(3600–3100 cm − 1), C=C and C=O stretching
(1740–1600 cm − 1),
C–O–C symmetric stretching (1097 cm − 1), –COOMe
(1400–1500 cm − 1), and so on

Table 1.

List of notable chemical characterisations of biochar.

Figure 7.

Biochar uses in agriculture and horticulture and its contribution to the circular economy [78].

ProcessProcess temperatureResidence timeSolid product yield on a dry wood feedstock basis (mass %)Carbon content of the solid product (mass %)Carbon yield (masscarbon, product/masscarbon feedstock
Slow pyrolysis~400Minute to days30950.58
Fast pyrolysis~500~1 s12–26740.2–0.26
Gasification~800~10–20 s10
HTC~180–2501-12 h<66%<70%0.88
Flash carbonization~300–600<30 min37850.65
Torrefaction~29010–60 min61–8451–550.67–0.85

Table 2.

Comparison of typical operating conditions and product properties of various biochar production processes [81].

5.1 Biochar in agriculture and horticulture

Biochar application in agriculture and horticulture has been explored both on a laboratory scale and in the field. These applications include being used as a component of chemical fertilizer [83], soil microbial activity, soil amendment for crop productivity improvement through nutrient availability [84, 85] as well as water holding capacity [86]. Biochar has also been reported to alleviate heavy metals release in the soil while having a limiting effect that aids in increasing the pH of highly acidic soils [87, 88]. Though biochar is another soil conditioner type, it differs from compost by production pathways. Biochar is produced by thermal decomposition of food, horticultural and municipal solid waste in the absence of oxygen, while natural biodegradation of organic substrates produces compost by the microbial community under aerobic conditions. Another difference is that; compost degrades fast, making its benefits relatively short-lived compared to biochar which persists in the soil for more prolonged periods [78, 89].

5.1.1 Biochar as a compost additive

Low soil organic carbon and fertility are challenges faced by many agricultural farmers around the globe. Biochar offers a solution to this challenge because it gives two options, i.e. returning nutrients and carbon to the soil while producing energy [90]. Also, the compositing rate can be increased by using biochar as an additive. Zhang and Sun [91] have examined spent mushroom compost and biochar co-composting. Their results showed a great increase in nutrients content of the resultant compost product and an improved composed quality while reducing the composting time from 90 to 270 days to only 24 days. Also, the large porosity of biochar enables it to facilitate microbial growth in the compost pile, leading to accelerated nutrient recycling [92]. The addition of biochar to poultry manure has been found to increase the maximum temperature reached and shorten the thermophilic phase [93].

5.1.2 Biochar as an adsorbent

An issue of heavy metals/metalloids (HMS) and polycyclic aromatic hydrocarbons (PAHs) in soil and water poses detrimental environmental problems and poor quality of agriculture, affecting all forms of life [94, 95]. These pollutants are toxic, persistent, non-biodegradable and potentially bioaccumulate [96]. Among other bioremediation technologies used to solve the HMS and PAHs issue, biochar is one of the best solutions due to its advantages [97]. These advantages include sustainability, low costs, sequestration of carbon, etc. [94]. Various physical and chemical characteristics of biochar, such as pore structure, specific surface area and functional groups, have been used to adsorption different pollutants [98]. For instance, Mahmoud, et al. [99] have used modified Switchgrass biochar for efficient decolorization of reactive red 195 A dye from aqueous and wastewater samples. Other biomass materials such as rice husks and dairy manure have also been used for biochar production with varying adsorption capacities according to the biomass used upon other factors [100].

5.2 Biochar in construction

Biochar has been used in road construction and as a concrete admixture. Wang, et al. [77] assessed this where a novel production of fill material and pedestrian/vehicle paving blocks were done. In this study, biochar addition was found to be beneficial to cement hydration even though it was noticed that the studied particle sizes could incur microcracks and strength degradation. Also, biochar’s incorporation resulted in enhanced immobilization of potentially organic contaminants and toxic elements in the sediment product, which is significant for moderately to heavily contaminated products. Therefore, biochar from wood can be used as a green combination for cement-based recycling procedures for highly contaminated waste. The use of biochar in construction material to trap atmospheric carbon dioxide in buildings also offers the potential to reduce greenhouse gasses by 25%. High pH and high water retention rate of biochar enable it to absorb some of the mixing water used in concrete mixing, thereby reducing the amount of free water in the concrete [101].

5.3 Future perspective

Since biochar’s applications depend greatly on its properties, future research must elucidate the production process effects on biochar’s properties. Biochar used in water treatment would differ from the one used in energy/agriculture. Likewise, there are diverse literature findings on the effects of biochar on agriculture, particularly on crop production caused by soils being different. For instance, crop yields may be increased or decreased by adding biochar depending on the soil type and fertilizer management [90, 102]. Also, the chemical behavior of biochar with heavy metal ions has been found to be inconsistent [103]. It is apparent that the interaction mechanisms between biochar, soil and plants are critical and yet not thoroughly known. Therefore, more efforts are still needed concerning biochar properties to soil and crop responses equally in the field and climate-controlled environment.

Advertisement

6. Conclusion

Biochar has been applied to remediate contaminated agricultural soil and improve soil fertility by reducing acidity and increasing the availability of nutrients. Thus, the addition of biochar to soils can be one of the best practices to overcome any biotic stress in soil and increase crop productivity, mainly in the agricultural sector. The properties of biochar have significantly been influenced by processes such as pyrolyscould, which have been discussed in this chapter. Thus, biochar appears as a highly promising option for pollutant removal. Economic impacts and recyclability should be considered in developing recoverable biochar for wide environmental applications. The relationship between various solutions for waste management and energy production differs in parameters and multiple techniques for its production and economic, social and ecological constraints. This review paper detailed the state-of-art information that would be helpful to find new opportunities in scientific innovation in the field of biochar research.

Advertisement

Acknowledgments

The authors are thankful to the Green Engineering and Sustainability research group in the Department of Chemical engineering at the Durban University of Technology, South Africa.

Advertisement

Conflict of interest

The authors declare no conflict of interest.

References

  1. 1. Amin FR, Huang Y, He Y, Zhang R, Liu G, Chen C. Biochar applications and modern techniques for characterization. Clean Technologies and Environmental Policy. 2016;18(5):1457-1473
  2. 2. Borel LD, Lira TS, Ribeiro JA, Ataíde CH, Barrozo MA. Pyrolysis of brewer’s spent grain: Kinetic study and products identification. Industrial Crops and Products. 2018;121:388-395
  3. 3. Borel LD, Reis Filho AM, Xavier TP, Lira TS, Barrozo MA. An investigation on the pyrolysis of the main residue of the brewing industry. Biomass and Bioenergy. 2020;140:105698
  4. 4. Zhou Y et al. Production and beneficial impact of biochar for environmental application: A comprehensive review. Bioresource Technology. 2021;337:125451
  5. 5. Yaashikaa P, Kumar PS, Varjani S, Saravanan A. A critical review on the biochar production techniques, characterization, stability and applications for circular bioeconomy. Biotechnology Reports. 2020;28:e00570
  6. 6. Al Arni S. Comparison of slow and fast pyrolysis for converting biomass into fuel. Renewable Energy. 2018;124:197-201
  7. 7. Bridgwater A, Peacocke G. Fast pyrolysis processes for biomass. Renewable and Sustainable Energy Reviews. 2000;4(1):1-73
  8. 8. Balogun AO, Sotoudehniakarani F, McDonald AG. Thermo-kinetic, spectroscopic study of brewer’s spent grains and characterisation of their pyrolysis products. Journal of Analytical and Applied Pyrolysis. 2017;127:8-16
  9. 9. Azargohar R, Nanda S, Dalai AK, Kozinski JA. Physico-chemistry of biochars produced through steam gasification and hydro-thermal gasification of canola hull and canola meal pellets. Biomass and Bioenergy. 2019;120:458-470
  10. 10. Wei J et al. Assessing the effect of pyrolysis temperature on the molecular properties and copper sorption capacity of a halophyte biochar. Environmental Pollution. 2019;251:56-65
  11. 11. Cantrell KB, Hunt PG, Uchimiya M, Novak JM, Ro KS. Impact of pyrolysis temperature and manure source on physicochemical characteristics of biochar. Bioresource Technology. 2012;107:419-428
  12. 12. Gabhane JW, Bhange VP, Patil PD, Bankar ST, Kumar S. Recent trends in biochar production methods and its application as a soil health conditioner: A review. SN Applied Sciences. 2020;2(7):1-21
  13. 13. Varma AK, Mondal P. Pyrolysis of pine needles: Effects of process parameters on products yield and analysis of products. Journal of Thermal Analysis and Calorimetry. 2018;131(3):2057-2072
  14. 14. Park HC, Lee B-K, Yoo HS, Choi HS. Influence of operating conditions for fast pyrolysis and pyrolysis oil production in a conical spouted-bed reactor. Chemical Engineering & Technology. 2019;42(12):2493-2504
  15. 15. 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
  16. 16. Liu W-J, Jiang H, Yu H-Q. Development of biochar-based functional materials: Toward a sustainable platform carbon material. Chemical Reviews. 2015;115(22):12251-12285
  17. 17. Li W et al. Mechanistic insight into lignin slow pyrolysis by linking pyrolysis chemistry and carbon material properties. ACS Sustainable Chemistry & Engineering. 2020;8(42):15843-15854
  18. 18. Xu W-C, Tomita A. The effects of temperature and residence time on the secondary reactions of volatiles from coal pyrolysis. Fuel Processing Technology. 1989;21(1):25-37
  19. 19. Yogalakshmi K et al. Lignocellulosic biomass-based pyrolysis: A comprehensive review. Chemosphere. 2022;286:131824
  20. 20. Chang C et al. Study on products characteristics from catalytic fast pyrolysis of biomass based on the effects of modified biochars. Energy. 2021;229:120818
  21. 21. Dai L et al. Production of bio-oil and biochar from soapstock via microwave-assisted co-catalytic fast pyrolysis. Bioresource Technology. 2017;225:1-8
  22. 22. Huang Y-F, Chiueh P-T, Kuan W-H, Lo S-L. Microwave pyrolysis of lignocellulosic biomass: Heating performance and reaction kinetics. Energy. 2016;100:137-144
  23. 23. Bridgwater AV. Renewable fuels and chemicals by thermal processing of biomass. Chemical Engineering Journal. 2003;91(2-3):87-102
  24. 24. Dabros TMH et al. Transportation fuels from biomass fast pyrolysis, catalytic hydrodeoxygenation, and catalytic fast hydropyrolysis. Progress in Energy and Combustion Science. 2018;68:268-309. DOI: 10.1016/j.pecs.2018.05.002
  25. 25. Demirbas A, Arin G. An overview of biomass pyrolysis. Energy Sources. 2002;24(5):471-482
  26. 26. Ghodake GS et al. Review on biomass feedstocks, pyrolysis mechanism and physicochemical properties of biochar: State-of-the-art framework to speed up vision of circular bioeconomy. Journal of Cleaner Production. 2021;297:126645
  27. 27. Tripathi M, Sahu J, Ganesan P, Dey T. Effect of temperature on dielectric properties and penetration depth of oil palm shell (OPS) and OPS char synthesized by microwave pyrolysis of OPS. Fuel. 2015;153:257-266
  28. 28. Tripathi M, Bhatnagar A, Mubarak NM, Sahu JN, Ganesan P. RSM optimization of microwave pyrolysis parameters to produce OPS char with high yield and large BET surface area. Fuel. 2020;277:118184
  29. 29. Lam SS et al. Microwave vacuum pyrolysis conversion of waste mushroom substrate into biochar for use as growth medium in mushroom cultivation. Journal of Chemical Technology & Biotechnology. 2019;94(5):1406-1415
  30. 30. Carrier M, Hugo T, Gorgens J, Knoetze H. Comparison of slow and vacuum pyrolysis of sugar cane bagasse. Journal of Analytical and Applied Pyrolysis. 2011;90(1):18-26
  31. 31. Chen N, Ren J, Ye Z, Xu Q , Liu J, Sun S. Study on vacuum pyrolysis of coffee industrial residue for bio-oil production. In: IOP Conference Series: Earth and Environmental Science. Vol. 59, no. 1. United Kingdom: IOP Publishing; 2017. p. 012065
  32. 32. Yazdani E, Hashemabadi SH, Taghizadeh A. Study of waste tire pyrolysis in a rotary kiln reactor in a wide range of pyrolysis temperature. Waste Management. 2019;85:195-201
  33. 33. Kebelmann K, Hornung A, Karsten U, Griffiths G. Intermediate pyrolysis and product identification by TGA and Py-GC/MS of green microalgae and their extracted protein and lipid components. Biomass and Bioenergy. 2013;49:38-48
  34. 34. Li G, Ji F, Bai X, Zhou Y, Dong R, Huang Z. Comparative study on thermal cracking characteristics and bio-oil production from different microalgae using Py-GC/MS. International Journal of Agricultural and Biological Engineering. 2019;12(1):208-213
  35. 35. Stummann MZ, Høj M, Gabrielsen J, Clausen LR, Jensen PA, Jensen AD. A perspective on catalytic hydropyrolysis of biomass. Renewable and Sustainable Energy Reviews. 2021;143:110960
  36. 36. Galiasso R, González Y, Lucena M. New inverted cyclone reactor for flash hydropyrolysis. Catalysis Today. 2014;220:186-197
  37. 37. Thangalazhy-Gopakumar S, Adhikari S, Gupta RB, Tu M, Taylor S. Production of hydrocarbon fuels from biomass using catalytic pyrolysis under helium and hydrogen environments. Bioresource Technology. 2011;102(12):6742-6749
  38. 38. Sekar M et al. A review on the pyrolysis of algal biomass for biochar and bio-oil–bottlenecks and scope. Fuel. 2021;283:119190
  39. 39. Melligan F, Hayes M, Kwapinski W, Leahy J. A study of hydrogen pressure during hydropyrolysis of Miscanthus x giganteus and online catalytic vapour upgrading with Ni on ZSM-5. Journal of Analytical and Applied Pyrolysis. 2013;103:369-377
  40. 40. Chen D, Gao A, Cen K, Zhang J, Cao X, Ma Z. Investigation of biomass torrefaction based on three major components: Hemicellulose, cellulose, and lignin. Energy Conversion and Management. 2018;169:228-237
  41. 41. Yu H, Zhang Z, Li Z, Chen D. Characteristics of tar formation during cellulose, hemicellulose and lignin gasification. Fuel. 2014;118:250-256. DOI: 10.1016/j.fuel.2013.10.080
  42. 42. Molton PM, Demmitt T. Reaction Mechanisms in Cellulose Pyrolysis: A Literature Review. United States. 1977.DOI: 10.2172/7298596
  43. 43. Lindstrom JK, Proano-Aviles J, Johnston PA, Peterson CA, Stansell JS, Brown RC. Competing reactions limit levoglucosan yield during fast pyrolysis of cellulose. Green Chemistry. 2019;21(1):178-186
  44. 44. Reza MS et al. Acacia holosericea: An invasive species for bio-char, bio-oil, and biogas production. Bioengineering. 2019;6(2):33
  45. 45. Yang X, Zhao Y, Li W, Li R, Wu Y. Unveiling the pyrolysis mechanisms of hemicellulose: Experimental and theoretical studies. Energy & Fuels. 2019;33(5):4352-4360
  46. 46. Huang J, Liu C, Tong H, Li W, Wu D. Theoretical studies on pyrolysis mechanism of xylopyranose. Computational and Theoretical Chemistry. 2012;1001:44-50
  47. 47. Mu W, Ben H, Ragauskas A, Deng Y. Lignin pyrolysis components and upgrading—Technology review. Bioenergy Research. 2013;6(4):1183-1204
  48. 48. Santos RB, Hart P, Jameel H, Chang H-M. Wood based lignin reactions important to the biorefinery and pulp and paper industries. BioResources. 2013;8(1):1456-1477
  49. 49. Brebu M, Vasile C. Thermal degradation of lignin—A review. Cellulose Chemistry & Technology. 2010;44(9):353
  50. 50. Ma Z et al. In-depth comparison of the physicochemical characteristics of bio-char derived from biomass pseudo components: Hemicellulose, cellulose, and lignin. Journal of Analytical and Applied Pyrolysis. 2019;140:195-204
  51. 51. Dasappa S, Shivapuji AM. Biomass gasification: Thermochemical route to energetic bio-chemicals. Advanced Biofuel Technologies: Elsevier. 2022;12:305-332
  52. 52. Galindo AL, Lora ES, Andrade RV, Giraldo SY, Jaén RL, Cobas VM. Biomass gasification in a downdraft gasifier with a two-stage air supply: Effect of operating conditions on gas quality. Biomass and Bioenergy. 2014;61:236-244
  53. 53. Prabakar D, Manimudi VT, Sampath S, Mahapatra DM, Rajendran K, Pugazhendhi A. Advanced biohydrogen production using pretreated industrial waste: Outlook and prospects. Renewable and Sustainable Energy Reviews. 2018;96:306-324
  54. 54. Maitlo G et al. Thermochemical conversion of biomass for syngas production: Current status and future trends. Sustainability. 2022;14(5):2596
  55. 55. Prins M, Ptasinski K, Janssen F. More efficient gasification via torrefaction. Energy. 2006;31:3458-3470. DOI: 10.1016/j.energy.2006.03.008
  56. 56. Sarker TR, Nanda S, Dalai AK, Meda V. A review of torrefaction technology for upgrading lignocellulosic biomass to solid biofuels. Bioenergy Research. 2021;14(2):645-669
  57. 57. Ruiz JA, Juárez M, Morales M, Muñoz P, Mendívil M. Biomass gasification for electricity generation: Review of current technology barriers. Renewable and Sustainable Energy Reviews. 2013;18:174-183
  58. 58. Bisht AS, Thakur N. Pine needles biomass gasification based electricity generation for Indian Himalayan region: Drivers and barriers. In: Green Buildings and Sustainable Engineering. New York, United States of America: Springer; 2020. pp. 47-59
  59. 59. Kivisaari T, Björnbom P, Sylwan C, Jacquinot B, Jansen D, de Groot A. The feasibility of a coal gasifier combined with a high-temperature fuel cell. Chemical Engineering Journal. 2004;100(1-3):167-180. DOI: 10.1016/j.cej.2003.12.005
  60. 60. Safarian S, Unnþórsson R, Richter C. A review of biomass gasification modelling. Renewable and Sustainable Energy Reviews. 2019;110:378-391. DOI: 10.1016/j.rser.2019.05.003
  61. 61. Basu P. Combustion and Gasification in Fluidized Beds. Florida, United States of America: CRC press; 2006
  62. 62. Puig-Arnavat M, Bruno JC, Coronas A. Review and analysis of biomass gasification models. Renewable and Sustainable Energy Reviews. 2010;14(9):2841-2851. DOI: 10.1016/j.rser.2010.07.030
  63. 63. Shahbeig H, Shafizadeh A, Rosen M, Sels B. Exergy sustainability analysis of biomass gasification: A critical review. Biofuel Research Journal. 2022;9(1):1592-1607
  64. 64. Islam MW. A review of dolomite catalyst for biomass gasification tar removal. Fuel. 2020;267:117095
  65. 65. Obemah ND, Zhao B. Biochar Preparation, Characterization, and Adsorptive Capacityand its Effect on Bioavailability of Contaminants: An Overview. China: Hindawi Publishing Corporation; 2014
  66. 66. Mary SG, Sugumaran P, Niveditha S, Ramalakshmi B, Ravichandran P, Seshadri S. Production, characterization and evaluation of biochar from pod (Pisum sativum), leaf (Brassica oleracea) and peel (Citrus sinensis) wastes. International Journal of Recyclying of Organic Waste Agriculture. 2017;5:43-53
  67. 67. Sarfaraz Q et al. Characterization and carbon mineralization of biochars produced from different animal manures and plant residues. Scientific Reports. 2020;10(1):1-9
  68. 68. Jindo K, Mizumoto H, Sawada Y, Sanchez-Monedero MA, Sonoki T. Physical and Chemical Characterization of Biochars Derived from Different Agricultural Residues. Spain: Biogeosciences; 2014
  69. 69. Laird DA. The charcoal vision: A win–win–win scenario for simultaneously producing bioenergy, permanently sequestering carbon, while improving soil and water quality. Agronomy Journal. 2008;100(1):178-181
  70. 70. Cha JS et al. Production and utilization of biochar: A review. Journal of Industrial and Engineering Chemistry. 2016;40:1-15
  71. 71. Figueredo NAD, Costa LMD, Melo LCA, Siebeneichlerd EA, Tronto J. Characterization of biochars from different sources and evaluation of release of nutrients and contaminants. Artigo Científico. 2017;48(3):395-403
  72. 72. Yang X, Ng W, Wong BSE, Baeg GH, Wang C-H, Ok YS. Characterization and ecotoxicological investigation of biochar produced via slow pyrolysis: Effect of feedstock composition and pyrolysis conditions. Journal of Hazardous Materials. 2019;365:178-185
  73. 73. Wang D, Jiang P, Zhang H, Yuan W. Biochar production and applications in agro and forestry systems: A review. Science of the Total Environment. 2020;723:137775
  74. 74. Woolf D, Amonette JE, Street-Perrott FA, Lehmann J, Joseph S. Sustainable biochar to mitigate global climate change. Nature Communications. 2010;1(1):1-9
  75. 75. Kah M, Sigmund G, Xiao F, Hofmann T. Sorption of ionizable and ionic organic compounds to biochar, activated carbon and other carbonaceous materials. Water Research. 2017;124:673-692
  76. 76. Pan J, Ma J, Zhai L, Luo T, Mei Z, Liu H. Achievements of biochar application for enhanced anaerobic digestion: A review. Bioresource Technology. 2019;292:122058
  77. 77. Wang L et al. The roles of biochar as green admixture for sediment-based construction products. Cement and Concrete Composites. 2019;104:103348
  78. 78. Hu Q et al. Biochar industry to circular economy. Science of the Total Environment. 2021;757:143820
  79. 79. Laird DA, Brown RC, Amonette JE, Lehmann J. Review of the pyrolysis platform for coproducing bio-oil and biochar. Biofuels, Bioproducts and Biorefining. 2009;3(5):547-562
  80. 80. Lehmann J. A handful of carbon. Nature. 2007;447(7141):143-144
  81. 81. Meyer S, Glaser B, Quicker P. Technical, economical, and climate-related aspects of biochar production technologies: A literature review. Environmental Science & Technology. 2011;45(22):9473-9483
  82. 82. Sohi SP, Krull E, Lopez-Capel E, Bol R. A review of biochar and its use and function in soil. Advances in Agronomy. 2010;105:47-82
  83. 83. Glaser B, Lehmann J, Zech W. Ameliorating physical and chemical properties of highly weathered soils in the tropics with charcoal–a review. Biology and Fertility of Soils. 2002;35(4):219-230
  84. 84. Pandit NR, Mulder J, Hale SE, Martinsen V, Schmidt HP, Cornelissen G. Biochar improves maize growth by alleviation of nutrient stress in a moderately acidic low-input Nepalese soil. Science of the Total Environment. 2018;625:1380-1389
  85. 85. Purakayastha T et al. A review on biochar modulated soil condition improvements and nutrient dynamics concerning crop yields: Pathways to climate change mitigation and global food security. Chemosphere. 2019;227:345-365
  86. 86. Mohamed BA, Ellis N, Kim CS, Bi X, Emam AE-R. Engineered biochar from microwave-assisted catalytic pyrolysis of switchgrass for increasing water-holding capacity and fertility of sandy soil. Science of the Total Environment. 2016;566:387-397
  87. 87. Rizwan M et al. Cadmium phytoremediation potential of brassica crop species: A review. Science of the Total Environment. 2018;631:1175-1191
  88. 88. Ruzickova J et al. A comprehensive assessment of potential hazard caused by organic compounds in biochar for agricultural use. Journal of Hazardous Materials. 2021;403:123644
  89. 89. Tratsch MVM, Ceretta CA, Silva LSD, Ferreira PAA, Brunetto G. Composition and mineralization of organic compost derived from composting of fruit and vegetable waste. Revista Ceres. 2019;66:307-315
  90. 90. Gaskin JW et al. Effect of peanut hull and pine chip biochar on soil nutrients, corn nutrient status, and yield. Agronomy Journal. 2010;102(2):623-633
  91. 91. Zhang L, Sun X. Changes in physical, chemical, and microbiological properties during the two-stage co-composting of green waste with spent mushroom compost and biochar. Bioresource Technology. 2014;171:274-284
  92. 92. Sanchez-Monedero M, Cayuela M, Roig A, Jindo K, Mondini C, Bolan N. Role of biochar as an additive in organic waste composting. Bioresource Technology. 2018;247:1155-1164
  93. 93. Czekała W, Malińska K, Cáceres R, Janczak D, Dach J, Lewicki A. Co-composting of poultry manure mixtures amended with biochar–the effect of biochar on temperature and C-CO2 emission. Bioresource Technology. 2016;200:921-927
  94. 94. Anae J et al. Recent advances in biochar engineering for soil contaminated with complex chemical mixtures: Remediation strategies and future perspectives. Science of the Total Environment. 2021;767:144351
  95. 95. Pérez RM, Cabrera G, Gómez J, Abalos A, Cantero D. Combined strategy for the precipitation of heavy metals and biodegradation of petroleum in industrial wastewaters. Journal of Hazardous Materials. 2010;182(1-3):896-902
  96. 96. Hadia-e-Fatima AA. Heavy metal pollution–a mini review. Journal of Bacteriol Mycology Open Access. 2018;6(3):179-181
  97. 97. Alaboudi KA, Ahmed B, Brodie G. Effect of biochar on Pb, Cd and Cr availability and maize growth in artificial contaminated soil. Annals of Agricultural Sciences. 2019;64(1):95-102
  98. 98. Beesley L, Moreno-Jiménez E, Gomez-Eyles JL. Effects of biochar and greenwaste compost amendments on mobility, bioavailability and toxicity of inorganic and organic contaminants in a multi-element polluted soil. Environmental Pollution. 2010;158(6):2282-2287
  99. 99. Mahmoud ME, Nabil GM, El-Mallah NM, Bassiouny HI, Kumar S, Abdel-Fattah TM. Kinetics, isotherm, and thermodynamic studies of the adsorption of reactive red 195 a dye from water by modified switchgrass biochar adsorbent. Journal of Industrial and Engineering Chemistry. 2016;37:156-167
  100. 100. Xu X, Cao X, Zhao L. Comparison of rice husk-and dairy manure-derived biochars for simultaneously removing heavy metals from aqueous solutions: Role of mineral components in biochars. Chemosphere. 2013;92(8):955-961
  101. 101. Gupta S, Kua HW. Factors determining the potential of biochar as a carbon capturing and sequestering construction material: Critical review. Journal of Materials in Civil Engineering. 2017;29(9):04017086
  102. 102. Asai H et al. Biochar amendment techniques for upland rice production in northern Laos: 1. Soil physical properties, leaf SPAD and grain yield. Field Crops Research. 2009;111(1-2):81-84
  103. 103. Abdelhafez AA, Li J, Abbas MH. “Feasibility of biochar manufactured from organic wastes on the stabilization of heavy metals in a metal smelter contaminated soil”. Chemosphere. 2014;117:66-71

Written By

Edward Kwaku Armah, Maggie Chetty, Jeremiah Adebisi Adedeji, Denzil Erwin Estrice, Boldwin Mutsvene, Nikita Singh and Zikhona Tshemese

Submitted: 19 April 2022 Reviewed: 26 April 2022 Published: 06 June 2022