Next Article in Journal
Performance Evaluation of a Virtual Test Model of the Frame-Type ROPS for Agricultural Tractors Using FEA
Next Article in Special Issue
Effects of Continuous Manure Application on the Microbial Community and Labile Organic Carbon Fractions
Previous Article in Journal
Agricultural Tractor Retail and Wholesale Residual Value Forecasting Model in Western Europe
Previous Article in Special Issue
Effects of Variations in Soil Moisture and Phosphorus Concentrations on the Diversity of the Arbuscular Mycorrhizal Fungi Community in an Agricultural Ecosystem
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Biochar Functions in Soil Depending on Feedstock and Pyrolyzation Properties with Particular Emphasis on Biological Properties

by
Polina Kuryntseva
1,*,
Kamalya Karamova
1,
Polina Galitskaya
1,
Svetlana Selivanovskaya
1 and
Gennady Evtugyn
2
1
Institute of Environmental Sciences, Kazan Federal University, Kazan 420008, Russia
2
Alexander Butlerov Institute of Chemistry, Kazan Federal University, Kazan 420008, Russia
*
Author to whom correspondence should be addressed.
Agriculture 2023, 13(10), 2003; https://doi.org/10.3390/agriculture13102003
Submission received: 8 August 2023 / Revised: 19 September 2023 / Accepted: 26 September 2023 / Published: 15 October 2023
(This article belongs to the Special Issue Ecological Environment and Microbial Community of Agricultural Soils)

Abstract

:
Biochar effects are strongly dependent on its properties. Biochar improves physical soil properties by decreasing bulk density and increasing medium and large aggregates, leading to faster and deeper water infiltration and root growth. Improvement of the chemical properties of soil is connected with pH neutralization of acidic soils, increase of cation exchange capacity and base saturation, providing a larger surface for sorption of toxicants and exchange of cations. Biochar increases the stocks of macro- and micronutrients in soil and remains sufficient for decades. Biochar effects on (micro)biological properties are mainly indirect, based on the improvements of habitat conditions for organisms, deeper root growth providing available C for larger soil volume, higher crop yield leading to more residues on and in the topsoil, better and deeper soil moisture, supply of all nutrients, and better aeration. Along with positive, negative effects of biochar while used as a soil conditioner are discussed in the review: presence of PAH, excessive amounts of K, Ca and Mg, declination of soil pH. In conclusion, despite the removal of C from the biological cycle by feedstock pyrolysis, the subsequent application of biochar into soil increases fertility and improves physical and chemical properties for root and microbial growth is a good amendment for low fertility soils. Proper use of biochar leads not only to an increase in crop yield but also to effective sequestration of carbon in the soil, which is important to consider when economically assessing its production. Further research should be aimed at assessing and developing methods for increasing the sequestration potential of biochar as fertilizer.

1. Introduction

In accordance with the International Biochar Initiative (IBI), biochar is defined as a solid material obtained by thermochemical carbonization in oxygen-limited conditions [1]. It is mostly produced during the pyrolysis process from organic matter at temperatures 300–700 °C at low access to oxygen [2]. Other pyrolysis products are liquid fuel and pyrolysis gas [3]. During the last decade, biochar has attracted attention in agroindustry and environmental sciences due to the prospects of its application as a soil conditioner and fertilizer, fodder additive, wastewater cleaning agent, renewable energy source and inexpensive sorbent used to remove heavy metals and organic pollutants. Biochar improves soil structure and fertility, promotes water holding capacity and is an important source of microelements and nutrients for plants, in stock and chicken farming [4]. Biochar application decreases methane and N2O emission [5,6], stimulates the activity of soil microorganisms and plant germination. It is important that biochar shows advantages by amendments in almost all agricultural conditions, including soil type and climate [7]. Contrary to many organic fertilizers, biochar carbon (C) is quite stable and is mineralized very slowly with the release of CO2 [8,9,10]. This makes biochar important in green sustainability policy, C sequestration, yield improvements and decrease of greenhouse gas emissions.
Biochar properties and hence the consequences of its application in agriculture depend on the biomass used for its production and the pyrolysis conditions [11]. Organic wastes including manure, plant residues, food wastes, sewage sludge, etc. have been used for soil fertilization since ancient times. In natural processes of biomass decomposition, about 90 to 95% of carbon is returned in the atmosphere within a few years (or earlier), whereas pyrolysis of bioresidues results in C sequestration and about half decrease CO2 emissions [12]. Therefore, biomass pyrolysis dramatically affects the C cycle (Figure 1). A slower rate of biochar mineralization prolongs its effect on soil fertility and decreases demands in other amendments. Biochar can also affect nitrogen cycling by mitigating N2O, another greenhouse gas also involved in atmospheric ozone decomposition, and by stimulating nitrification via indirect influence on the soil microbial community [13].

2. Biochar Production, Main Sources and Properties

2.1. Feedstock Materials

Biochar is commonly produced by pyrolysis of plant biomass, animal and food residues, forestry wastes, crop residues (nut shells, fruit pits, bagasse, straw [14,15,16,17]), algae [18,19], manures [11,20,21,22], biosolids and sewage sludge formed in wastewater treatment, etc. [23,24,25,26]. In the latter case, the content of heavy metals should be limited to avoid possible contamination of the soil. The elemental content of biochar (mainly, C, N, K, P, Ca and Mg) depends on the feedstock source and pyrolysis parameters (Table 1).
The use of animal manures often results in increased potassium content. Generally, mineral components of biochar vary from 1% (grasses) to about 20% (wastewater sludge). C content of the biomass used as biochar feedstock depends on the source, pre-treatment conditions, storage period, moisture content and other parameters, some of which can be partially considered prior to thermal treatment for unification of the pyrolysis parameters and product characteristics. The classification of biomass applied for biochar production is mainly based on its source, biological diversity and origin [27].
Plant residues, especially forest and grass residues, mainly contain cellulose, lignin and hemicellulose (lignocellulosic biomass (Dhyani and Bhaskar 2018) [28]). Cellulose is present in all plant cell walls. It consists of repeating D-glucose units interacting with each other by hydrogen bonds, providing mechanical strength and chemical stability. Hemicellulose is a less polymerized analog of cellulose containing glucose, galactose, mannose, xylose, arabinose and glucuronic acid units (50–200 monomers per molecule). Lignin is a cross-linked phenol polymer consisting of differently substituted phenylpropane units. Lignin forms the outer layer of cellulose fibers and binds hemicellulose and cellulose within the cell wall. Lignin content varies from 16 (soft wood) to 40% (hard wood) (Table 2).
In addition to the three main components, organic matter of biomass contains some extractives (alkaloids, essential oils, glycosides, pectin, phenolics, simple sugars, terpenes etc.) They can be removed from biomass by organic solvents without hydrolysis. Details of the biopolymers’ chemical structure and their physical properties as well as examples of the ash content of the biomass and ash are presented in review [41].
Organic matter of biosolids (more than 50% of dry matter) mostly contains hydrocarbons, amino acids and lipids but a rather small percentage of lignin and cellulose [26]. Lignin and cellulose content is higher in urban wastes but remains significantly below the level typical for wood or crop residues. Biosolids are also rich in nitrogen, mostly present in organic form (up to 6% of dry weight) and phosphorus (up to 3% of dry weight), but they can be partially lost in combustion and pyrolysis (the loss of P is less than the loss of N). Biosolids themselves and biochar produced from biosolids are characterized by a high ash content, which complicates the use of this parameter for calculations of the biochar yield.
The classification of biochar feedstocks mostly uses moisture content as one of the key features. Freshly cropped biomass, including agricultural plant and forest residues, sewage sludge, algae, animal manure, etc., typically contains more than 30% of water (Table 3). They are referred to as wet biomass. Agricultural wastes with a water content below 30% are classified as dry feedstock. Wet biomass can be pre-dried using technologies developed for drying other biomaterials. However, in most cases, such technologies are labor consuming and decrease the total economic effect of biochar application [42]. Biomass is classified on that specially harvested as bioenergy crops (bamboo, sorghum, willows, Miscanthus, switchgrass [43,44] and agricultural wastes [28]. The use of energy crops provides a high yield of the target product and minimal preliminary treatment is required. Thus, no pre-drying is commonly needed. The ash content of such a biomass can significantly vary depending on the harvesting time and planting density complicates the selection of optimal pyrolysis parameters and feedstock unification. Biomass from various wastes is more variable regarding its content. This type of feedstock involves wastes formed in agriculture, forestry, food production, sewage sludge from water treatment plants, animal manure, etc. Their conversion to biochar is always beneficial because it reduces expenses for their environmentally safe storage and/or processing. On the other hand, parts of such wastes that do not contain toxic metals and organic contaminants can be applied on land as soil amendments and organic fertilizers with no or minimal treatment. Thus, full processing of the plant biomass into biochar seems exhausting. It should also be taken into account that crop residue removal negatively affects the soil organic carbon pool, which should be reimbursed by agrochemical measures [45]. Preliminary treatment of biomass is one of the key factors affecting both the pyrolysis products and the repeatability of the technological parameters of biochar production. Moisture content, preliminary crushing or, vice versa, pressing of the feedstock are most frequently utilized on the preliminary stage of pyrolysis [46]. Appropriate protocols influence both the economy and preferable directions of biochar application.

2.2. Pyrolysis Conditions

The technological process of biomass pyrolysis starts with supplementary drying of the substrate if its moisture is higher than optimal. The following heating removes the natural volatile components present in the raw material. After that, decomposition of the biopolymers present in biomass results in consecutive increases in the C content and release of permanent gases (CO2 and nitrogen oxides, ammonia and methane) and volatile organic compounds of decomposition. The latter can be condensed and then used as liquid fuel. The solid residue is biochar. The reaction pathways of the formation of gaseous and liquid pyrolysis products partially compete with each other and differ depending on the pyrolysis parameters (residence time, heating rate, pyrolysis temperature) and substrate properties. Thus, fast heating increases the yield of pyrolysis gases and fuel, whereas slow and rather long pyrolysis is more appropriate for predominant carbonization and maximum biochar yield.
Thermal treatment of biomass can be performed using several regimes, which are chosen depending on the type of feedstock, its pre-treatment, desired properties and future application of biochar. They are classified as slow, fast, flash and intermediate pyrolysis, dry torrefaction and hydrothermal carbonization, which are performed in the absence or limited access of oxygen. Pyrolysis can also be accelerated by microwave treatment. The characteristics of pyrolysis reported for some feedstock and regimes are presented in Table 4 as an example.
Long pyrolysis is a traditional technology of biochar production with heating at 300–700 °C by the rate of 5 to 20 °C/min for more than one hour. In traditional biochar production, carbonization can take place for more than one week. It produces a maximum yield of biochar against other technologies mentioned above (more than 50% of dry weight). Hard wood can be heated up to 1000 °C with the final carbonization level. Meanwhile, biomass from agricultural waste and other sources with rather mild temperature of ash melting is normally heated at not more than 700 °C [61,62].
Fast pyrolysis is mostly used for the production of liquid fuel: volatile products are condensed by rapid cooling to avoid polymerization of by-products. The biomass is heated for several seconds to pyrolysis temperatures. Gases contain various quantities of carbon oxides, methane and hydrogen [55]. Fast pyrolysis may result in incomplete conversion of biomass so that up to 9% of hydrocarbons (mostly in the form of cellulosic and hemicellulosic fractions) remain in the product. Their presence diminishes later C sequestration in soil. Raising the pyrolysis temperature decreases the content of this fraction but due to lower amounts of biochar obtained.
Flash pyrolysis [56] is performed by ignition of the press biomass at elevated pressure (1–2 MPa). The reactions provide heating of the feedstock up to 300 to 600 °C for about 30 min. The yield of biochar increases with temperature and pressure [57].
Dry torrefaction is a process in which biomass is heated in an inert atmosphere at 200 to 300 °C for the period of 30 min to 2–3 h. Up to 30% of biomass is lost, including a 10% decrease in energy content. Polysaccharides are depolymerized to solid residues with a low O/C ratio. Torrefaction assumes slow heating and can be categorized as mild pyrolysis. Torrefaction is commonly used for pre-treatment of biomass prior to its combustion or gasification. The product of dry torrefaction still contains significant amounts of volatile organic matter from the feedstock.
Gasification is partial combustion of the biomass at a high temperature (600 to 1200 °C) for several seconds [63]. As follows from the name of the process, gas consisting mostly of methane and hydrogen is formed. Up to 10% of the dry weight of the biomass is converted into biochar. Unlike other biochar sources, it can be contaminated with polyaromatic hydrocarbons and heavy metals so that it cannot be recommended for application as a soil amendment. Nevertheless, its application improves the physical properties of the soil and positively affects microbial activity [58].
Hydrothermal carbonization is performed at 180 to 260 °C with biomass dispersed in water under elevated pressure (2–6 MPa) for 5 to 240 min [64,65]. The char produced has a higher carbon content than the products of dry pyrolysis [59]. Its characteristics depend on the reaction temperature, pressure, residence time and biomass/water ratio. An increase in the temperature results in a higher yield of liquid products. At 250 °C, up to 70% (dry weight) of biochar are obtained. Preliminary treatment of the feedstock with mineral acids and salts increases the yield and decreases optimal temperature and pressure due to better solubilization of the biomass and suppression of hydrogen bonds between the biopolymer molecules [60,66].
Chemical reactions during biomass pyrolysis depend on substrate composition and heating conditions. Cellulose, hemicellulose and lignin comprising biomass undergo various conversion paths, including depolymerization, cross-linking and decomposition of the fragments. Hemicellulose decomposes at 200 to 260 °C, cellulose breaks at 240 to 350 °C and lignin starts decomposing at 280 to 500 °C [67]. Slow pyrolysis at rather low temperatures results in decomposition and carbonization of cellulose, whereas fast pyrolysis promotes volatilization and formation of levoglucosan [68], which is later decomposed by C-O, C-C scission and dehydration to low-molecular compounds [58,69]. In a similar manner, hemicellulose undergoes cleavage to 1,4-anhydro-D-xylopyranose yielding furfural and 2C-, 3C-fragments [70]. Lignin is involved in drying and fast and slow degradation steps result in the formation of a number of compounds, e.g., phenolics [63]. Regarding other elements, sodium and chlorine are released at a relatively low temperature, whereas calcium and magnesium can be bonded to organic species and removed at higher temperatures [71]. Alkali and alkali-earth metals exert autocatalytic effects on biomass pyrolysis due to the acceleration of secondary destruction of volatiles produced in pyrolysis and formation of more gases affecting biochar cracking [64]. Phosphorus, sulfur and nitrogen are mostly bonded in less complex organic compounds present in the plant cells and decomposed in slow pyrolysis at rather low temperatures. What becomes of various chemical elements in biomass during its pyrolysis and mechanisms of chemical reactions responsible for the formation of organic species, biofuel and biochar is described in detail in a review [65].

2.3. Improving Biochar Properties by Modification Approaches

Biochar application often assumes pre-treatment aimed at introduction of additional functional groups, increase of porosity and specific surface area [72]. Such a treatment is also desirable for the application of biochar for catalyst production because it increases the adsorptive capacity of potential catalytic sites, such as transient metal cations. Meanwhile, biochar application may have unexpected and undesired effects, e.g., formation and accumulation of toxic species (polyaromatic hydrocarbons). The general classification of biochar pre-treatment approaches is presented in Figure 2.
The introduction of additional functional groups starts from partial oxidation of biochar. Hydrogen peroxide [72], ozone [73], potassium permanganate [74] (and nitric acid [75] are used for this purpose. Oxidation products contain carboxyl, phenolic, peroxide fragments and lactones. Oxidation increases the hydrophilicity of the char surface. Similar results can be obtained by heating biochar in an inert atmosphere with low amounts of oxygen. The appropriate changes depend on the treatment temperature. Thus, post-treatment of chars obtained from carbohydrates and activated by heating to 300 °C in the presence of oxygen was investigated by FT-IR spectroscopy and X-ray fluorescence analysis [76]. Treatment below 450 °C promotes opening inner space without mesopore widening and results in preferable surface adsorption of inorganic components. Treatment temperature increase to 900 °C removes most oxygen-containing functional groups and stimulates the formation of more common aromatic structures.
Oxidation within the range of 450 to 900 °C makes mesopores wider and deeper. The contribution of the adsorption of inorganic species on initial organic matter decreases. Metal cations remain attached in pores and on the surface due to electrostatic interactions with aromatic π-electron systems. Meanwhile, full pore volume, surface area and capacity toward anions (phosphates) can decrease during high temperature pyrolysis [58,69]. High temperature post-treatment decreases the biochar mass by about 10% per each 100 °C. In turn, this increases the yield of mineral residue and the content of metal oxides and carbonates [77].
In addition to oxidation in the presence of small amounts of oxygen, micropore widening is achieved by controlled biochar gasification. For this purpose, it is treated with water vapor or carbon dioxide. This increases the internal volume of micropores, but the number of mesopores increases insignificantly. Thus, post-activation of biochar from coconut shells has been described with various rates of vapor and carbon dioxide supply with simultaneous microwave treatment [78]. The specific surface square was increased to 2000 m2/g, and micropore volume exceeded 80% of the total internal volume of the activated product. Simultaneous microwave treatment and water vapor/CO2 treatment increased the product yield twofold [79].
Some catalysts, i.e., zinc chloride and phosphoric acid, make for porosity increase and are added to raw biomass prior to pyrolysis. Being absorbed by biomass, they prevent gumming in the first heating stages. Phosphoric acid hydrolysis glycoside bonds in hemicellulose and cellulose and breaks ether bonds in lignin. The reactions mentioned are often followed by degradation and dehydration of the cellulose structure and condensation of oligomeric by-products. In such a conversion, carbon oxides and methane are released. Phosphoric acid can also form esteric bonds with hydroxide groups of biomasses if temperatures do not exceed 200 °C. This promotes the cross-binding of polymeric chains [80]. As a result, carbon retention, sorption capacity and pore formation are synchronously improved [81]. Unlike phosphoric acid, zinc chloride acts as a dehydration agent. It decreases carbonization temperature for cellulose, hemicellulose and lignin and changes paths of their following chemical conversion in favor of gumming suppression and open pore formation. ZnCl2 can then stimulate wood swelling due to its implementation in the biomass and depolymerization processes. Zinc chloride remains in melted conditions up to the end of pyrolysis (melting point below 700 °C) and hence inhibits C rearrangement in the structures to be obtained. Removal of ZnCl2 from biochar leaves well-pronounced micropores. Unlike other chemical agents, zinc chloride maximizes the increase of the pore volume in biochar. Other additives, e.g., FeCl3, mostly affect biomass graphitization and the formation of graphitic particles with a high specific surface area (up to 2000 m2/g) [82].
Biochar amination is the second most widespread method of biochar functionalization. It simplifies CO2 capture and toxicant sorption [83,84]. Amination starts with preliminary biochar oxidation, followed by its ammonia addition [85,86]. Polyethyleneimine [87,88] is involved in similar reactions. In addition, amino groups are formed by nitration of biochar followed by reduction of nitro groups introduced with dithionite [62]. Aminated biochar finds applications in binding metal cations. In addition, amino groups change the hydrophilicity of the surface and improve biochar wettability, which is important for its field application. It should be mentioned that in addition to chemical modification, increased nitrogen content is achieved by an environmentally safer approach based on pyrolysis of biomass enriched with nitrogen-containing materials, e.g., chitin [89] and chitosan [90]. Such modifiers also increase sorption of metal cations [91] and are applied in the preparation of composite materials with the implementation of magnetic particles [92].
Biochar sulfurization (introduction of sulfonate groups) can be achieved by post-treatment of biochar with sulfuric acid, filtration and centrifugation of target products [93]. They are used for sorptional removal of toxic metals and as solid acids in soil remediation and wastewater treatment [79,94].
Most of the modified biochar particles have negative surface charges associated with a few carboxylic, carbonyl and sulfonate groups. Therefore, they can adsorb metal cations that are additionally bonded by complexation with amino (amide) groups. High ash content in the substrate increases the accumulation of metals, whereas higher pyrolysis temperatures affect this parameter in the opposite way due to partial destruction of acidic groups. Instead, the products of high temperature pyrolysis adsorb increased quantities of organic species, e.g., catechols or humic acids [80]. This is referred to as the general increase of the pore volume [95]. However, they do not effectively bind anions, e.g., nitrates, arsenates or phosphates. To avoid this limitation, metal oxides can be introduced into the biochar structure. The feedstock is soaked in a solution of metal nitrates or chlorides. Their following heating in the presence of atmospheric oxygen to about 300 °C releases nitrogen oxides (chlorine) so that metals are fixed in the biochar material in oxide form. Meanwhile, impregnation of metal oxides regularly decreases the biochar surface area due to partial filling of the pores [96]. Treatment of the biochar obtained from the corn comb by slow pyrolysis at 500 °C with Fe(III) nitrate resulted in 20 times higher accumulation of As [97]. Hybrid material obtained by pyrolysis of the mixture of naturally occurring hematite (γ-Al2O3) and pinewood biomass showed a much stronger ability to remove arsenic from wastewaters than biochar with no impregnation [98]. A similar efficiency was registered for biochar impregnated with gelatin containing ferrite magnetic particles [99]. Introduction of AlO(OH) [100], manganese oxides [101] and MgO [102] was also described for soil remediation and toxic metals removal from the sediments and wastewaters. It is interesting to mention that biochar containing MgO was obtained from tomato tissue biomass that accumulated Mg from irrigation waters [103].
In addition to metal oxides, chemical modification of biochar with clay minerals, e.g., kaolinite, montmorillonite and bentonite, has been described as increasing accumulation of ammonia and phosphates. The product obtained by mixing biomass with appropriate materials prior to pyrolysis is enriched with Al, Fe and Na [104,105]. Bentonite keeps its structure in the final product, whereas montmorillonite is dehydrated. Maximum sorption capacities of biochar for ammonia and phosphates exceeded 12.5 and 105 mg/g, respectively.
Biochars were also mixed with carbonaceous nanomaterials, e.g., graphene oxide [106] and carbon nanotubes [107]. They increase the porosity of the product, specific surface area and surface concentration of oxygenated functional groups. As a result, they accumulated metal cations better than the biochars themselves. Hybrid materials described have been utilized for the removal of Cr(VI) and Hg(II) from the wastes of sugar production.

2.4. Biochar Properties

2.4.1. Physical Properties

The physical properties of biochar are pre-determined by the processes taking place during carbonization. Biochar obtained in low temperature pyrolysis consists of graphitic layers. The distance between the layers, as well as their specific surface area, increases with the carbonization temperature [108]. Aromatic carbon atoms dominate the biochar structure obtained at 350 °C with rather small contribution of alkyl oxygen and C atoms. An increase in the pyrolysis temperature to 500 °C converts alkyl fragments into aryl groups with a low H/C ratio. Hydrothermal conversion results in the formation of roundish particles. The model study showed that they resulted from the preliminary formation of lignin-like structures with a high number of oxygen-containing functional groups (ether, quinone and pyron groups) distributed within hydrophobic internal layers covered by hydrophilic coats [46]. The nature and distribution of the surface functional groups were established using FTIR spectroscopy [109] and XRD analysis [110]. Biochar obtained at low temperatures contains a shell formed by furan-like pentatonic rings (60% of carbon atoms) bridged with sp2- (sp3-) C atoms [111]. At higher temperatures, such layers are converted into condensed aromatic rings to form graphitic particles. Based on solid-phase 13C NMR, FT-IR spectroscopy and GC-MS, the conversion starts at 270 °C and is accompanied by the evolution of volatile furan derivatives [112]. Similar behavior has been reported in other works utilizing various feedstocks and pyrolysis regimes [77,113,114].
The biochar particle morphology results from the chemical reactions during pyrolysis. At low temperature (at about 300 °C) free radicals appear [115]. The following heating destroys cellulose with the formation of anhydrosugars that are less reactive than radical products of the bond cleavage. Being volatile, such intermediates are condensed in the pores of biochar and then in secondary carbonization form a typical crystal lamellar structure [116]. The surface groups of biochar depend on the pyrolysis rate. Oxygen containing hydroxide and carboxylic groups are mostly present on the materials obtained in fast pyrolysis, whereas aromatic C-H bonds are mostly found on the surface of slow pyrolysis products [117]. Heating changes ratio of main components and structural properties (Figure 3).
The structure of the biochar particles changes slightly with the heating rate but is sensitive to the final pyrolysis temperature. With its increase, crystal fragments become bigger and their internal structure—more regular [118]. The variety of the structure fragments is higher in fast pyrolysis. As gases are released from the solid matrix, bulk density decreases. This results in a lower bulk density that also takes into account free space between the biochar particles. Thus, pyrolysis of woody biomass decreases density twofold against feedstock at the pyrolysis temperature of 350 °C, which is considered a lower limit of fastest changes of biochar porosity [119]. Similar changes in porosity were observed for grass biomass heated at 350 to 700 °C [120]. For the plant substrate rich in hemicellulose, the temperature promoting pore formation can be decreased due to the lower thermal stability of hemicellulose. The porosity changes caused by lignin decomposition at higher temperatures are more complex. On the one hand, decomposition of lignin can result in even higher porosity than that of hemicellulose carbonization. On the other hand, the shrinking of solid residue influences this parameter in the opposite direction [121].
It should be mentioned that the lower density of biochar against substrate is related to bulk density. The so-called true density describing solid phase disregarding pore space increases with residence time and pyrolysis temperature. Shrinking biomass during pyrolysis can also result in incomplete removal of volatiles, especially in low temperature pyrolysis (Figure 4) [122].
Changes in the specific surface area during pyrolysis follow porosity increase. Thus, it starts raising at 400–500 °C until 900 °C from about 10 to 100–500 m2/g depending on the feedstock and the pyrolysis conditions [123,124]. The following heating above 900 °C leaves surface area constant or slowly decreasing. This was attributed to changes in the particle structure resulting in pore widening or collapse of some pore walls. In addition, decreased area can be influenced by ash melting and fusing [125]. Similar consequences are expected from secondary reactions between char and volatiles remaining entrapped in the inner pore volume [126] and from crystallization of amorphous carbon into graphite [127]. Thus, the average pore size of the biochar particles obtained from sewage sludge increased from 4.7 nm (feedstock) to 8 and 28 nm at 500 and 600 °C and then decreased to 16 nm at 700 °C [128]. The surface area of the wood pyrolysis products obtained at 450 °C increased from 4 to 23 m2/g with the residence time varied from 10 to 60 min, respectively [19]. Biochar from rubber wood sawdust increased the surface area from 1.93 to 5.49 m2/g with the pyrolysis temperature raised from 300 to 700 °C [129].
Figure 4. Molecular structure of biochar from plant biomass across a charring temperature gradient; (A) Characterization of organic phases. (B) Char composition from gravimetric analysis [130]. With permission of the American Chemical Society.
Figure 4. Molecular structure of biochar from plant biomass across a charring temperature gradient; (A) Characterization of organic phases. (B) Char composition from gravimetric analysis [130]. With permission of the American Chemical Society.
Agriculture 13 02003 g004
In addition to porosity, the surface square depends on the particle size distribution [131]. The average size of the biochar particles regularly decreases with the pyrolysis temperature due to weakened macromolecular structure and formation of fragile particles likely to be broken [132]. Thus, pitch pine pyrolysis increased the fraction of particles below 500 μm from 70 to 95% (pyrolysis at 300 °C) [133].

2.4.2. Chemical Properties

The chemical composition of biochar is mostly determined by carbonization resulting from the loss of hydrogen and oxygen containing groups. The progress of carbonization is reflected by changes in the atomic rations. Regarding three main elements, i.e., carbon, oxygen and hydrogen, such changes are expressed by a diagram presented first by D. van Kleveren [134] (Figure 5). In natural carbonization there are various natural ways of carbonization. However, the release of oxygen is twofold faster than that of hydrogen until coal is formed. After that, the decrease in the H/C ratio becomes constant if the oxygen content is maintained at a low level. In technical conditions, the relative decrease of the O/C and H/C molar ratios can vary, but contrary to natural conditions, the appropriate rates do not change within the pyrolysis duration. The van-Kleveren diagrams are present in many reviews and original works devoted to biochar production. A summary of 212 experimental results obtained for various pyrolysis conditions and biomass origins is presented in [135]. Maximum decrease of the atomic ratios corresponds to a 250 to 350 °C interval in which the O/C and H/C ratios decrease twofold. The following removal of oxygen requires heating to 700 °C whereas full removal of hydrogen requires even higher heating. Van Kleveren diagrams reflect a general assessment of biochars and related products from the point of view of energy storage and energy losses in combustion.
Carbon, hydrogen, oxygen and minor quantities of nitrogen are the main components of biochar. Elemental composition follows changes in the atomic ratios corresponding to the appropriate temperature of pyrolysis and heating rate. In addition, moisture, silicon, phosphorus and metal residues should be considered. The particular content of biochar highly depends on the substrate used. Carbon content varies commonly from 45 to 60 wt.%, hydrogen from 2 to 5 wt.% and oxygen at the level of 10 to 20% [136]. Inorganic components (minerals) are present in a higher amount in plant resides and algae, and in a lower amount in woody biomass. Biochar obtained from plants contains more carbon than that from manure (Table 5).
In most cases, carbonization of biomass promotes increase of the pH of the product to final pH = 8–10. The higher the pyrolysis temperature, the more the pH shift. In a similar manner, pH correlates with ash content. Thus, the use of woody biomass results in the formation of the product with a pH lower than that of manure carbonization products rich in mineral components. The consideration of the pH is complicated by the changes in acidity observed within the time after implementation of biochar in soil. As an example, the pH of the biochar obtained from oak chips decreased within a year of application to soil, whereas the coals from corn wastes increased from 6.7 to 8.1 [138]. Such relationships are related to the destruction of acidic functional groups at high temperatures of pyrolysis. Another factor influencing the chemical composition of biochar is particle size. A number of authors have found that different particle sizes of raw materials can lead to different heating during pyrolysis, so a smaller fraction could be more carbonated than a larger fraction and contain more C, as well as Ca, Mg and K ions [131,139].

3. Biochar Functions in Soil

Biochar is commonly used in rehabilitating infertile or degraded soils because it improves soil physicochemical characteristics, reduces greenhouse gas emissions, increases fertilizer application efficiency, increases yields, and absorbs organic pollutants [140]. Biochar affects physical and chemical soil properties directly, and biological properties mainly indirectly. The direct effects are mainly connected with feedstock sources and pyrolysis conditions.

3.1. Effects on Soil Physical Properties

The physical properties of biochars (microstructure of biochar particles) are crucial for their effects on soils. The biochar bulk density varies from 1.5 to 2 g/cm3 [117] and that of soil particles varies from 2.4 to 2.8 g/cm3. The application of 30 t/ha of woody biochar reduced the soil bulk density by 14% on a fallow land [141]. The bulk density decreased linearly with the biochar application rate in the range from 0 to 100% vol. from 2.62 to 1.60 g/cm3 [142]. In most cases, linear dependence of the biochar amendment and soil density decrease were observed with maximum effect for application of about 60 t/ha and no influence of small application rates (<10 t/ha) [143]. In coarse textured solids, reduction of the bulk density caused by biochar is higher than coalescence that in fine textured soils. The largest decrease, by 31%, reported in 2016 was found for sand [144]. This is explained by the mixing effect due to the rather high difference in the above parameters of the soils and chars used.
The mean size of the soil aggregates used for estimation of many important soil properties, e.g., water filtration, macropore development, soil particle cracking, etc., is also expressed as the mean size (diameter) of the aggregates present in amended soil. Biochar increases wet aggregate stability with no respect to the soil type and application rate due to the ability of organic matter to establish bonds between soil particles. The release of binding molecules as well as direct aggregation of biochar particles with organomineral complexes results in the formation of aggregates and their following coalescence. The effect is more pronounced in sandy soils [145,146]. Aggregate size increases have also been observed in grassland and volcanic soils with organic carbon levels of >10% [147,148]. Clay mineralogy is another factor altering the biochar influence on soil aggregates. Similar relationships were found for dry aggregation stability that requires a rather high application rate (20–25 t/ha) of biochar and was preferable on sandy soils [149,150].
Changes in soil bulk density affect soil porosity. This is mostly referred to as the soil C content, which varies with biochar amendment and hence influences soil porosity [151]. The porosity is normally increased due to the combination of various factors, e.g., reduced bulk density, increase of the aggregation, changes in the interaction with mineral components, and denser packing of the soil particles. For example, fine-textured soil (clay) is less affected by bio-char than coarse-textured soil (sandy) in terms of the degree of change in bulk density. To reliably reduce soil density, the application rate of biochar should vary in the range of 1–5% [152]. It has been shown that biochar can increase soil porosity by up to 19% and reduce bulk density by up to 35% [153]. To increase the overall porosity of the soil and reduce its bulk density, it is necessary to use a highly porous biochar, i.e., biochar obtained at low and medium temperatures and obtained from plant raw materials. Thus, biochar obtained by wheat bran pyrolysis at 800 °C increased soil porosity to a lower extent than that obtained at 1200 °C [154].
Tensile strength refers to the inherent capability of the soil to resist external forces that can cause fractures and ruptures. Biochar application on clayey soils decreases tensile strength because clays have viscoelasticity exceeding that of biochar. The tensile strength decreased from 64 to 32 kPa after 50 t/ha of biochar application and to 19 kPa for 100 t/ha [155]. In sandy soils, however, biochar added alone or together with organic fertilizers improves soil structure and increases tensile strength due to better aggregation [156].

3.2. Effects on Soil Chemical Properties

Great attention has been paid to studying the use of biochar for the remediation and restoration of polluted soils [157,158,159,160]. Its beneficial effects are most often attributed to immobilizing toxicants and reducing the bioavailability of toxins to soil organisms. An increase in the pyrolysis temperature increases the carbonization degree of biomasses and the surface of biochar particles, which contributes to an increase in their sorption capacity for organic pollutants and hence a decrease in their potential hazard to soils. In particular, a decrease in the bioaccumulation of polychlorinated biphenyls by soil organisms was found when activated carbons were introduced into soils, supported by an adequate change in the equilibrium concentration of the toxicant in the study of their extraction from water systems [161,162]. A similar effect of biochar on the equilibrium concentrations of organic toxicants was established for DDT [163] and polycyclic aromatic compounds [164]. Sorption of organic toxicants does not preclude their subsequent release from the bound form over time. Biochar also binds heavy metals in a carbon matrix. In particular, a relationship has been registered between a decrease in the mobility of copper, cadmium and nickel during the period of decomposition of biochar introduced into the soil and the content of organic matter in it [165]. Therefore, its production and application to soils can be considered an effective strategy in limiting the circulation of such elements in the environment. A fairly large number of examples of biochar use with a view to suppressing the phytotoxic properties of heavy metals have been described. For example, a regular decrease in the phytotoxic concentrations of cadmium and zinc was registered in a 60-day experiment with biochars introduced into silt masses [166,167]. The amounts of retained heavy metals correlate with the cation-exchange capacity of biochar, which provides evidence of the mechanism of their retention on the carboxylate groups of biochars. At the same time, an inverse relationship was found for copper, associated with an increase in soluble organic carbon in soil moisture when biochar from hardwood was introduced. This proves that the specific effects of biochar application on the bioavailability of toxic soil elements require verification for actual application conditions. There is a necessity to certify the maximum allowable levels of toxic elements in biochar [168]. The ability of biochar to remove arsenic from wastewater has been found [169]. Unlike metals, arsenic is present in soil in the form of oxo anions. Accordingly, the mobility of arsenic in soils does not decrease with pH increase, as in the case of copper, zinc, lead and cadmium, but increases. Its retention in soils is due to the presence of aluminum, iron and manganese oxides in them, which react with arsenic forms. Thus, there is a possibility of increasing the bioavailability of this element due to the fact that the introduction of biochar increases the pH and soluble carbon of soils. In laboratory tests aimed at assessing the influence of biochar on the arsenic content in wastewater moderately polluted with heavy metals, the effect of biochar introduced in an amount of 30 vol.% was not reliably detected [170]. Additional benefits in the removal of arsenic can be expected from biochar enriched with iron oxides, and if used in conjunction with compost, a simultaneous decrease in the bioavailability of other common heavy metals can be achieved [171]. Similar patterns were revealed for the binding of chromium (VI) compounds.
There are works devoted to the use of biochar produced from chicken manure with the aim of removing metals from the environment. Thus, its high efficiency with respect to copper (II) ions has been proven. The share of the removed metal increases with biochar mineralization and varies from 1.3 to 26 mg/g of the sorbent. The use of chicken manure as a raw material provided a higher capacity of biochars compared to coals produced from corn waste, pine wood, and eucalyptus [172]. A decrease in copper bioavailability was also observed when chicken manure biochar was applied to copper-contaminated sandy loamy soils. Biochar was prepared by slow pyrolysis at 500 °C for 2 h [173]. It was characterized by a high content of phosphorus (19.4 g/kg) and potassium (17.2 g/kg), pH 9.1 (water suspension 1:5) and a specific surface area of 11.5 m2/g. The interaction of biochars with copper was also controlled by changing the growth rate of O. picenis; the indicator species is a metallophile. In model experiments in seed germination in a greenhouse, biochar was applied in the amount of 5 and 10%. In addition to the effect on plants, a decrease in the concentration of copper in soil moisture and an increase in soil microbial activity are described. The manifestation of the phytotoxic properties of biochar with a high content of a number of other trace elements (phosphorus, boron, zinc, manganese) in relation to plants sensitive to their presence in the soil was registered [174,175]. There are other risks associated with the agronomic use of biochar. Thus, it was found that the introduction of 1% biochar made of eucalyptus chips halved the bioavailability of two insecticides (chlorpyrifos and carbofuran) [176]. Sorption of pesticides can help in reducing crop contamination, but it reduces the efficacy of their application against insect pests. In addition, the biochar itself may contain phytotoxic compounds [177]. The influence of biochar on the content of polycyclic aromatic hydrocarbons (PAHs) is the focus, as they are partially formed in the pyrolysis of organic matter, the precursor of biochar. Adsorption or other protection against microbiological degradation of PAHs limits their removal from polluted soils; saturation of biochar pores with other adsorbed substances reduces its effect on the removal of those compounds into groundwater. It is not entirely clear how the influence of biochar on the microbiological community and its protective effect on PAHs correlate, as both adsorption and microbiological activity depend in a complex way on soil pH after the application of biochars. Examples of agrochemical and ecotoxicological consequences of biochar application to soils are given in Table 6 [178].
In general, an increase in the pyrolysis temperature during the production of biochar increases the sorption efficiency in relation to organic pollutants. Presumably, this occurs as a result of an increase in the specific surface area and microporosity of coals.

3.3. Effects on Soil Biological Properties

Although the positive effect of biochar on soil fertility and agriculture ecosystems mainly refers to the pH changes in acidic soils and improved nutrient retention and aeration, remarkable influence of biochar amendments on microbial communities and soil animals has been reported [186,187,188]. Changes in microbiota composition and abundance also follow soil structure improvement and nutrient availability caused by biochar application (indirect biochar effect). All of these factors influence plant growth and crop quality and consequently, more input of rhizodeposits into the soil. The positive influence of biochar amendments can be counterbalanced by toxic metals contained in the pyrolysis products, especially those derived from sewage sludge or wastewater sediments. The consideration of the biochar effects on soil microorganisms reported in the literature mainly depends on what kind of effect is expected, i.e., biodiversity, abundance, or secondary effects on soil biota on crops and soil fertility. Thus, soil biota diversity is often believed to be a key factor for soil functions [189] and organic amendments derived from biochar are key factors affecting trophic relations in soil [190]. The second approach assumes indirect influence of biochar on nutrient availability and possible effects of various species adsorbed in biochar particles and released in the environment during biochar aging and mineralization. Microbial abundance always increases with biochar amendments with low dependence on the feedstock and soil type. The mycorrhizal response estimated by root colonization increased twice for a two-year exposition of 0.6–6 t/ha of biochar to Eucalyptus wood [191]. The protective effect of biochar particles, sorption of signaling compounds, detoxification of allelochemicals, and improvement of physical properties of soil were mentioned as possible reasons for the positive effect of biochar on extraradical mycelium [3]. In addition, biochar was reported to be able to stimulate spore germination [192]. An opposite effect of biochar amendment was attributed to symbiotic relations in conditions of high nutrient availability, undesirable pH shifts and toxic effects of metal cations or high salt content [193,194]. Some other effects of biochar on the microbial abundance affected by its application are summarized in Table 7.
Fertilizers commonly reduce the influence of biochar application on microbial abundance. The effect is mostly pronounced for P, N-containing fertilizers that inhibit acceleration of root colonization observed with biochar added alone [195]. The plants did not rely on biological N fixation in the presence of fertilizers as much as in the presence of biochar. Non-symbiotic organisms can slightly increase their abundance in the same conditions due to the higher availability of nutrients from fertilizer. Biological nitrogen fixation can also be stimulated by introduction with soybean straw-derived biochar [196]. Microbial biomass increases with pH values caused by biochar in acidic soils. The effect is more pronounced for bacteria, whereas fungi are more tolerant to this factor and can even reduce their growth at high pH [197]. It should be taken into account that changes in the soil pH observed after biochar application can reflect secondary effects on microbial activity [192]. The pH effect significantly depends on the soil properties. Biochars added to soil protect microorganisms from the inhibitory effect of some species (catechols, flavonoids, phenolics) that are adsorbed on the pores and hence excluded from the microorganism metabolism. Such an effect was described for biochar obtained from corn stover [198]. Fast-pyrolysis biochar from wood promoted arbuscular colonization of asparagus due to limitations in access to some aromatic acids exerting an allelopathic effect.
The microbial response to the biochar addition is also sensitive to the adhesion of the cells on the particle surface. This formation of biofilms mostly affects bacterial but not fungal abundance. The attachment of the cells to the biochar particles can occur via hydrophobic attraction and electrostatic forces. The mineralization of the biochar, as well as higher salt content, accelerate adhesion. Inclusion of the cells into the pores is limited by their size, which should be at least 2–5 times bigger than the cell size. For Bacillus mucilaginous and Acinetobacter sp., it was estimated of 204 μm [150]. In addition to abundance, microbial community composition significantly differs with biochar addition against that of unmodified soils of the same mineralogy [151]. Viscosity decreases stimulate the development of plant roots and rhizobia because of the higher availability of nutrients. Additionally, invertebrates move easier through the amended soil, making pores and influencing predator/prey ratio. A similar influence on plant roots and fungi can be attributed to the changes in the soil bulk density caused by the rather small value of this parameter of biochar particles (typically between 0.09 and 0.5 g/cm3 against 1.5–2.1 g/cm3 for true density).
Transformation of soil nutrients. It has been established that nitrification is accelerated when biochar is added to forest soils, which is explained by the sorption of phenols that inhibit the process and an increase in the mass of ammonium-oxidizing bacteria [199]. In addition, an increase in the activity of alkaline phosphatase, aminopeptidase and N-acetylglucosamine oxidase was revealed upon the introduction of biochar [200], which may be associated with an increase in the production of organic nitrogen and phosphorus due to the accelerated growth of plant roots in the pores of biochar. Bradyrhizobiaceae (Rhodoblastus, Rhodopseudomonas, Bradyrhizobium and Nitrobacter) may take part in the process, as well as nitrogen utilizing Hyphomicrobiaceae (Rhodoplanes, Starkeya), nitrates and ammonia in nitrogen fixation or denitrification [201]. The same microorganisms can produce ethylene from fresh biochar, which in turn leads to a reduction in greenhouse gas (N2O and CO2) emissions [202]. To study nitrogen fixation by legumes (Phaseolus vulgaris L.) in soils treated with 0–90 g/kg biochar soil, the isotope dilution method (15NH4)2SO4 was used [203]. At the maximum dose of biochar, the proportion of fixed nitrogen increased from 50% in the control to 72%. The increase in total nitrogen supplied from the atmosphere was 49% for the dose of 30 g/kg of biochar and 78% for 60 g/kg, a decrease was registered at the maximum dose of application. A possible reason for the increase in microbial nitrogen fixation was an increase in the availability of boron and molybdenum and, to a lesser extent, potassium, calcium and phosphate, an increase in pH, a decrease in the availability of nitrogen, and saturation of aluminum. At the same time, the input of soil nitrogen decreased by 14 and 17% at an application rate of 30 and 60 g/kg of biochar, respectively. Biochar has a positive effect on phosphorus-mobilizing mycorrhiza, as it protects mycelium, indirectly influencing the changes in the physicochemical characteristics of soils, and detoxifying allelopathic secretions [204,205]. There are also opposite observations, indicating the absence of the influence of biochar or its negative effect on mycelium. One explanation for this contradiction is the indirect effect of biochar due to changes in pH, saturation of the soil with oxygen, and changes in its porosity.
Microbial diversity in soils enriched with biochar has been studied using various methods, including total DNA analysis of the soil microbial community [206], colony culture and counting [207], substrate-induced respiration [208,209], microbial biomass [210,211], extraction of phospholipid fatty acids [212,213,214,215], contrasting, and direct examination of individual biochar particles [3]. The introduction of biochar increases microbial diversity in different ways for different groups of microorganisms. The two most common types of mycorrhizal fungi (arbuscular and ectomycorrhizal fungi) most often respond positively to the application of biochar (see review [3]). The response of mycorrhiza is usually assessed by root colonization. Thus, the rate of formation and the number of processes of ectomycorrhizae on larch seedlings in the presence of biochar in the soil increased by 20–160% [216]. Similar results were obtained on wheat in a two-year experiment with the introduction of biochar from eucalyptus wood (0.6–6.0 t/ha)—the acceleration of colonization was 5–20% relative to the control without the introduction of biochar [191]. It is less clear how the part of the mycelium located in the soil interacts with the biochar. The direct interaction of biochar and mycelium may be important. For example, the internal pores of the biochar particles can protect the extraradial mycelium from external influences, for example, from grazing animals or from soil overconsolidation [188].
The introduction of biochar at doses of 5 and 25 t/ha was studied when cultivating wheat for 10 weeks while simultaneously varying the dose of applied nitrogen [217]. It was shown that Cmic decreased when biochar was introduced, while Nmic remained practically unchanged. The experiment cannot be explained by the sorption of inorganic nitrogen on coals, since carbon dioxide emissions decreased in the presence of 5 t/ha of biochar, but not in the presence of 25 t/ha. The structure of the ammonium-oxidizing microbial community changed only when biochar was introduced together with a nitrogen source. The authors concluded that the introduction of biochar reduced the activity of the microbial community as a whole. An increase in the dose of mineral fertilizers reduces the positive effect of biochar on the rate of reproduction of microorganisms [218], depending on the nature of the fertilizer and the group of microorganisms. Thus, mycorrhizal infection is suppressed by phosphorus-containing fertilizers, regardless of the application of biochar, but it does not depend on the application of nitrogen fertilizers. The opposite situation was registered for Rhizobium [219]. The reason may be the different influences of external conditions on the symbiosis of microorganisms. Therefore, when nitrogen fertilizers are added, the plant may not need biological nitrogen fixation to the same extent as in the absence of top dressing. A similar reason may explain the different effects of increased carbon load in the rhizosphere during exudation. The microbial diversity of non-symbiotic microorganisms can increase with increased availability of nutrients, either as a result of their longer retention in soil enriched with biochar or due to their entry into soils together with biochar. In most cases, a similar improvement can be achieved by the direct application of nitrogen and phosphorus fertilizers without biochar.
An increase in the concentration of microelements, such as molybdenum or boron, can also cause an increase in biological nitrogen fixation by Rhizobium colonizing biochar [220]. This phenomenon was discovered during the study of the immobilization of toxic elements by biochar. Similarly, sorption reduces the toxic effect of antimicrobial compounds and elements by reducing the time of their action, especially when they first enter the soil [221].
An example of the indirect effect of biochar on soil microorganisms is a change in pH. An increase in microbial biomass (Cmic) from 20 to 180 μg/g and in ninhydrin nitrogen from 0.5 to 4.5 μg/g was registered with an increase in soil pH from 3.7 to 8.3 [222]. At the same time, bacteria and fungi react differently to changes in pH. Bacteria increase diversity with an increase in pH to 7 or more, while fungi do not change their biomass in the specified pH range and decrease biomass with further increase [197,223]. Therefore, biochar with pH varying within the specified range can affect soil biota differently depending on soil pH. In addition, even with the introduction of acidic samples, for example, hydrochar with pH 4, an increase in the pH of fertilized soils was observed, apparently due to secondary processes of microbial reduction of unidentified organic substrates and electron transfer mediators [192]. The said work indicates the importance of secondary processes in fertilized soils, which, in turn, are determined by the pH value of the soil before its cultivation, the direction and magnitude of the change in pH when biochar is added, etc. In addition, it should be taken into account that the pH on the surface of biochar particles, where a bacterial film is formed, may differ from that in the bulk of the soil [188].
In addition to changing the microbial biomass, the introduction of biochar and related changes in the conditions of microbiological community development change its structure up to the change of dominants. Such changes have been repeatedly recorded for Terra preta (artificial soil based on low-grade activated carbon, also called Brazilian black soil), soils enriched with coal from natural fires and soils with biochar and for soil fungi, bacteria and archaea [207,224,225,226]. As a rule, the diversity of the bacterial community increased up to 25% when biochar was applied to the soils of Terra preta but decreased with a similar tillage in soils after natural fires. Simultaneously, in these cases, the diversity of archaea and fungi decreased, which indicates the unequal effect of biochar on various components of soil communities. However, it should be noted that these studies were carried out at different times, from 0.5 and 2.5 years for soils after fires and with the introduction of biochar and up to hundreds and thousands of years for Terra preta soils.
The bacterial community in soils with a high content of biochar differs from that in soils possessing the same mineralogy without biochar in them [207,227]. This was also shown with the help of genetic fingerprinting of Terra Preta soils and unmodified forest soil in the Amazonian region. The first ones showed a larger number of unique taxonomic units. At the same time, all taxa present in the forest soil were found in Terra preta, while the latter contained unique units that were absent in the forest soil. The greatest difference between the microbiological communities of both soils was established for an evolutionary distance of 5%, which indicates the presence of differences at the genome and species levels. Studies [207] revealed a high taxonomic diversity in biochar-fertilized Terra Preta soils in four areas of the central Amazonian region compared to unmodified soils. The maximum difference is fixed at the family level. Similar discrepancy values of 80% were observed for soils enriched with biochar and agricultural soils differing in cultivation time, past and present crops, and other conditions [206]. For uncultivated soils, the diversity did not exceed 40%. Up to the present, the biochar of Terra preta soils, introduced hundreds and thousands of years ago, remains the main factor determining the development of the microbiological community, despite the dissimilarity of other factors, such as agricultural use, soil texture, mineralogy, nutrient content and pH. Although the application of fertilizers, especially for podzolic soils, leads to differences in the results of genetic fingerprinting, it is incomparably small with the effect of biochar additions, which at times drastically changed the microbiological community. An increase in the application of biochar to the soils of the temperate zone increased the divergence of the bacterial community composition in the rhizosphere and soil. Rhizospheric soils with the application of 12 and 30 g/ha of biochar were characterized by the greatest dissimilarity from the soil, with minimal or no additional fertilization (0 or 1 g/ha) [188]. It was concluded that the introduction of biochar into soils leads to the development of communities similar to rhizospheric ones, which were formed without the introduction of biochar.
Similar studies were carried out at the level of taxa. Thus, two new taxa of Acidobacteria [227], and later two potentially new taxa of α-Proteobacteria were identified in the soils of Terra preta. It was shown that Acidobacteria are widely represented in all studied soils and differ in their genetic profile from analogs included in databases by at least 2% of the code. Some genomes isolated from Brazilian soils enriched with biochar were grouped with 93% similarity with Verrucomicrobia, whose genome is mainly found in tropical rice husks, but is characterized by a gradual spread to other soils. Genomes assigned to Pseudomonas, Acidobacteria and Flexibacter sp. were found both in Terra preta soils and in control soils. In moderately rich soils with and without biochar additions, 70% of the isolated sequences are classified as Ascomycota, Basidiomycota or Zygomycota. However, the occurrence frequency of the main phylotype genes differed for soils enriched and not enriched with biochar. In the presence of biochar, the communities were characterized by less genetic diversity. Similarly, less diversity was found in the Archaean community in Terra Preta soils, particularly in the ammonium-oxidizing Chrenarcheota [225]. Probably, this effect indirectly reflects changes in soil pH caused by the introduction of biochar [228]. Soils enriched with biochar show multiple Zygomycota involved in the degradation of glucose and cellulose and forming Glomeromycota mycorrhiza. At the same time, the diversity of Basidiomycota decreases by a third compared to unenriched soils. Some Ascomycota are known to degrade lignin but can also use simpler organic compounds as a substrate. In this regard, the lack of available carbon in biochar can inhibit the colonization of this fungal species, while soluble carbon adsorbed on the surface of biochar particles will increase the proportion of Zygomycota that apparently find sufficient resources of oxidizable carbon.
Similarly, under the influence of biochar, the bacterial community also changes. In response to the introduction of biochar obtained in high temperature pyrolysis (pyrolysis products of oak wood and a mixture of herbs at 650 °C), bacterial diversity increases, including taxonomy, in contrast to the effect of biochar obtained at 250 °C [226]. An increase in the diversity of Actinobacteria and Gemmatimonadetes was found, in agreement with similar observations of Terra preta soils and forest soils naturally fertilized with fire coal [229]. Separately, the effect of biochar on pathogenic microorganisms should be mentioned. The application of biochar to soils can increase E. coli contamination of groundwater, especially since current approaches do not consider reducing microbiological diversity in soil moisture as a goal. The transfer of E. coli to soil and ground moisture depends on the biochar application rate and the pyrolysis temperature. Thus, high temperature biochar (700 °C) from poultry litter at a dose of 2% applied to sandy soils does not reduce the number of E. coli in the leachate; increasing the dose to 10% (not realistic in real land use for economic reasons) reduces the number of microorganisms by several orders of magnitude [230]. Soil treatment with the same biochar, produced at 350 °C, increases the abundance of E. coli itself, which should be taken into account when assessing the consequences of its use in agriculture.
Along with the conservation of carbon and the reduction of the greenhouse effect in general, the introduction of biochar into soils has a general soil-improving effect, mainly associated with agronomic and environmental factors. There are various assumptions about the nature of such an effect, but most of them associate biochar with improved storage of soil moisture and nutrients, improved soil structure and drainage [231]. There is evidence of a connection between the application of biochar and the state of the soil microbiological community. These factors also affect the yield of agricultural plants, sometimes indirectly, for example, through the acceleration of nitrogen fixation by free and symbiotic diazotrophs [232]. The ability of biochars to absorb pollutants should also be taken into account, thereby reducing their availability for biota during the restoration of damaged soils and the neutralization of production wastes [203]. At the same time, such processes can lead to secondary contamination of soils with the same pollutants in the future during the decomposition of biochars.
Biochar stability in soil, i.e., its resistance to degradation, leaching and chemical oxidation, depends on the aromatic structure, surface functional groups, and sorption characteristics for minerals and organic compounds [233]. Biochar destruction proceeds under the influence of biotic factors (microbial community) due to photooxidation and dispersion [234]. Observations of freshly prepared biochar during the year showed that a decrease in its mass is accompanied by an increase in the number of surface oxygen atoms in the composition of carboxyl and phenolic groups, a decrease in the positive charge of the surface and the acquisition of a negative charge by it [138]. The O/C ratio determines the stability of a biochar, as expressed by its half-life, from >1000 years (O/C < 0.2) to <100 years at O/C > 0.6 [181]. It is difficult to accurately determine the lifetime of biochar in the soil, since the inhomogeneity of samples and the tendency of biochar to modify depending on environmental conditions increase during exposure.

3.4. Mineralization of Biochar

Biochar has a high sequestration potential due to the conversion of easily accessible C from plant biomass, organic waste, etc., into a hard-to-reach form, due to the absence of a priming effect when used as a soil improver, due to non-energy-intensive production (the resulting pyrolysis fuel and pyrolysis gas ensures the maintenance of the process, without the use of additional fuel sources), as well as by increasing the frequency of application (every 3–10 years) compared to other organic fertilizers, such as compost or digestate [22,235,236,237,238,239,240,241,242]. Evaluation of retention time of biochar carbon in soils varies from hundreds to thousands of years, while biomass carbon is retained for several decades [243]. The introduction of biochar derived from biomass reduces the return of carbon to the atmosphere in the form of carbon dioxide. Thus, provided that the gases released in parallel, as well as the products released into the ecosphere during the production, transportation and storage of biochar, do not compensate for the positive effect, the introduction of biochar contributes to a decrease in the greenhouse effect [244]. The manifestation of such an effect depends on the degree of carbon conservation in the biochar, i.e., on the ratio of carbon content in biomass and the same value in biochar obtained from the specified amount of biomass. Slow pyrolysis gives a carbon conservation value of about 50%; higher values are typical for a less stable biochar with a stable existence time in soils of 4 to 29 years [245].
The rate of natural mineralization of coals under natural conditions is usually low. Its evaluation is difficult because the changes over a reasonable period of observations are too small and because of the difficulty of unambiguous identification of the chemical nature of the resulting products. One of the methods of such an assessment is the use of biomass enriched in the 14C carbon isotope in the production of biochar. In particular, it was experimentally possible to prove the mineralization of 6% biochar, obtained in this way from a chaff, to CO2, during an 8.5-year experiment. This corresponds to the processing of 0.3% of the carbon introduced per year [246]. In another laboratory experiment with the carbonization products of 14C-labeled Pyrenean chaff, the rate of mineralization was estimated from the rate of 14CO2 release after the introduction of coals into loess soils [247]. It amounted to no more than 0.5% of the introduced carbon per year, which, in terms of real conditions, gives a lifetime of coals in soils of at least 2000 years. At the same time, the joint application of glucose to intensify the co-metabolism of soil bacteria reduces this time by more than 10 times.
The literature provides a fairly wide range of data on the carbon footprint of biochar, from 0.04 tCO2-eq to a net reduction of 3.9 tCO2-eq per ton of raw material [153,248,249,250,251,252]. This wide range is obtained due to different pyrolysis feedstocks, different pyrolysis plants, and different system boundaries established during the biochar life cycle assessment. A similarly wide range is set for the price of biochar. According to calculations by Nematian [253] and co-authors, the cost of producing 1 ton of biochar in the USA is USD 449–1847, in European countries EUR 300–2000 [254].Calculations made for Russia showed that the cost of production of 1 ton of biochar is USD 40; at a cost of USD 110/t, biochar production becomes profitable, which is comparable to other soil fertilizers and immobilization agents. It is worth noting that the growing demand for biochar, according to the European Biochar Market Report 2022/2023 biochar production capacity continues to grow in 2022 by 52% to 53,000 t biochar [252]. Indeed, the cost of compost in European countries and the USA is about 10 times lower than that of biochar and amounts to USD20–50/t. However, it is important to consider that when producing 1 ton of biochar in European countries, up to 4 tons of carbon credits worth EUR 70–370/t can be returned [153,254,255]. In addition, despite the same application doses, the frequency of applying biochar is once every 2–5 years, in contrast to the annual application of compost. In connection with the above, the most rational thing is to calculate the carbon footprint of crop products grown using biochar, as well as to change its cost taking into account the carbon units obtained.

3.5. Biochar Influence on Soil Microorganisms

Biochar influences many soil processes, such as denitrification, methane oxidation [256,257], carbon mineralization [210,247] and the transformation of nutrients. The reasons for this are numerous. They include switching to other carbon sources, changing nutrient availability, sorption of inorganic and organic components, including enzymes, to biochar, soil moisture retention, and changes in infiltration or pore structure. Later, those that are directly or indirectly related to the microbial community of soils will be considered.
As mentioned above, the microbial community is able to actively respond to the introduction of biochar. Higher microbial diversity could potentially lead to greater mineralization or oxidation of the biochar itself, as shown for organic carbons of non-pyrolysis origin. Usually, these processes are stimulated by an increase in microbial biomass. However, some reports have indicated the opposite effect or a combination of decreased diversity and absolute respiration carbon (carbon turnover) [210,258]. This may be the consequence of the lower amount of available carbon, the high stability of biochar, or the adsorption of organic carbon on it, which could contribute to its slower degradation. Vice versa, the introduction of freshly prepared biochar from waste products usually increases both respiratory activity and community metabolism [208]. It is explained, inter alia, by a high content of nutrients, such as nitrogen and phosphorus, as well as a significant proportion of organic biochar, which, in particular, follows from a slight activation of respiration with increasing biochar application rates [259,260]. In the same work, a positive connection was described between the amount of volatile organic components in the biochar and carbon dioxide released during incubation. Both of these factors, i.e., an increase in the nutrient component and labile carbon, directly follow the introduction of biochar into soils, and their overall effect on biochar mineralization depends on the ratio of labile carbon and nutrient components both in the biochar itself and in the soil inorganic matter.
From the same standpoint, it would be logical to expect a multidirectional effect of biochar on the structure of the microbial community. The shift toward a greater variety of fungi after the introduction of biochar into the soil potentially indicates a greater mineralization of the biochar itself. Root fungi are known to promote the degradation of lignin in woody biomasses and coal [261,262]. Interestingly, among fungi, there is a shift in diversity toward taxa that prefer glucose as a source of carbon, and among bacteria, in the opposite direction. It is not entirely clear how the introduction of a much more oxidatively tolerant biochar could contribute to the development of such preferences, especially given the wide variety of other carbon sources available in soils (own organic matter, litter, etc.). It is possible that mineralization of biochar does not increase access to labile carbon but promotes the mineralization of available non-pyrolysis carbon. This statement is consistent with observations showing a connection between an increase in microbial biomass and a higher rate of soil carbon decomposition (so-called priming) in the presence of biochar. The fact that this increase usually remains below the initial higher salinity when freshly prepared biochar is added [209] suggests that there are other mechanisms leading to carbon loss, such as physical carbon removal, changes in nutrient content or pH. In addition, volatile carbon compounds present in biochar, along with similar smoke components, can stimulate microbial activity immediately after the application of biochar to soils, but then quickly mineralize [263,264]. Longer-term (more than a year) observations show that biochar reduces the rate of soil carbon mineralization [137,210,265]. A similar situation (more biomass but less soil respiration) was also registered in liquid waste treatment, when the biofilm on irrigated fields showed a higher rate of mineralization of soluble aromatic carbon than the same film on activated carbons with a larger specific surface area of the latter [266]. It is possible that carbon dioxide forms carbonates on the surface of the biochars due to the higher pH value of their surface. This explains the decrease in the amount of recorded CO2 with a simultaneous increase in microbial biomass.
Changes in the composition and enzymatic activity of the microbial community upon the introduction of biochar into soils may be responsible for the lower mineralization of soil carbon. The activity of glycosidase and cellobiosidase decreased with the introduction of more than 12 t/ha of biochar. In addition to soil analysis, a similar change was recorded for the purified enzyme and biochar obtained by rapid pyrolysis from Panicum virgatum [200]. In such experiments, it is also necessary to take into account the localization of microorganism colonies on biochar particles. Their proximity to a carbon source increases the efficiency of carbon mineralization, even without additional enzyme production. An alternative explanation associated with the assumption of adsorption of enzymes on biochar seems less probable; for example, lipases form stable and highly effective adducts with activated carbons [267]. Biochar particles can generate regions of local enrichment in available carbon and thereby promote the growth of microbial colonies, as occurs in biological wastewater treatment [268]. In the latter case, immobilization of biochars also reduces the possible toxic effect of organic waste on microorganisms, which contributes to their higher metabolic activity [269]. Biochar can adsorb large amounts of soil organic carbon, which was demonstrated by studying microbial cultures and the processes of substance leaching from soil horizons [270]. Numerous evidence of strong adsorption of aromatic hydrocarbons on various kinds of coals and soot also proves this statement [271]. Although such processes, with the inclusion of microbial biomass and litter, are slower than the sorption of PAHs, they can significantly determine the influence of coals on the mineralization of soil carbon. The influence of biochar on multicellular soil organisms is of great interest from the point of view of monitoring the state of the microbiological community based on the following considerations: soil organisms are included in the general flow of soil matter and energy, they are present at the upper levels of the food chain, and their reactions are derived from the reactions of microorganisms to the introduction of biochar. Then, geophages, for example, earthworms, are important participants in the processes of biochar modification and its transfer to other soil horizons, which cannot but affect microorganisms. Finally, the reactions of soil organisms to toxic elements may reflect corresponding changes in the structure and abundance of the microbiological community.
The interaction of biochars with earthworms has been the most studied. They are able to absorb particles of coal, crushing them along the way and mixing them with the soil. In a number of experiments with microcosms, it was shown that worms prefer mixtures of soils with biochar over pure soils [180]. Perhaps they use coal particles to ground soil organic matter, or they are attracted to microorganisms and their metabolic products that accumulate in the pores of the biochar. Long-term plot experiments have shown that worms also contribute to the transportation of biochar particles within the soil layer but most likely not beyond it. At the same time, in similar experiments with PAH-contaminated biochar, a decrease in the mass of worms was observed in comparison with the inhabitants of the same soils without biochar [272]. The joint enrichment of soils with biochar and worms increased the concentrations of inorganic nitrogen, as well as the productivity of agricultural crops (for example, rice [273]). Data on the influence of biochar on nematodes are scarce. Soils treated with smoke from the production of activated carbons show a higher number of nematodes, which may be due not to the direct action of the components contained in the pyrolysis products but to their effect on competitive interaction in soils. The same can be said about arthropods, the change in the abundance of which in soils enriched with coals was established from the products of their vital activity.

3.6. Effect of Biochar on Crop Yields

Biochar application usually improves crop productivity by increasing the availability of nutrients, increasing the activity of microorganisms that determine the mobility of soil nutrients [274,275,276,277], and accelerating the development of the plant root system because of colonization of arbuscular mycorrhizal fungi and improved soil physical properties [278]. There is also a decrease in the loss of nutrients due to their reduced leaching from soils and an improvement in the physical characteristics of soils that contribute to water retention and oxygenation [279,280]. There is also data on the absence of the biochar effect or even its negative effect on crop yields. In Table 8, there is an example of generalization of the biochar effect on crop yields. Table 8 also contains information on the application rate and statistical parameters of the assessment [281,282].
When evaluating the contribution of biochar to soil fertility, it is necessary to take into account all aspects of a complex interrelated system that includes, in addition to biochar and soil, a specific agricultural crop, inorganic and organic fertilizers used in parallel, biochar application volumes, climatic and other environmental conditions (Table 8).
In Table 9 the results of testing various types of biochar in a field experiment are given. The effect of biochar on crop yield depends on soil type. For example, corn showed a 2–3-fold increase in yield when applied at 4 t/ha in acid sandy soils and only 30% to 40% when biochars were used in sandy-clay acid soils. In neutral clayey soils, no effect of biochar application was found.
In the study [295], a 50% increase in the yield of chaff was established on slightly acidic fatty loams and 44% on calcareous sandy soils. In Australian calcareous soils, a negative effect of biochar on the yield of lettuce and hay grass was registered [296]. Technological aspects complicate the dependence of yield on the type of soil, cultivated crops and the method of biochar preparation. Thus, in a greenhouse experiment, a decrease in the yield of corn in the first year of cultivation was registered when biochar from eucalyptus wood obtained in fast pyrolysis at 800 °C was introduced into acidic marl soils [297]. At the same time, biochar did not affect this value when applied to weakly acid silty loams. When using biochar obtained in slow pyrolysis at 350 °C, no increase in yield was observed for either of the two soil types. Different results were obtained in the second year of the experiment: after the introduction of biochar obtained in slow pyrolysis, the yield increased by 500% on slightly acidic loams and by 150% on acidic marl soils. The authors attributed the reasons for the decrease in the efficiency of using the fast pyrolysis product to the possible presence of PAHs and the antagonistic effect of excess potassium on calcium and magnesium, although reliable results for measuring these chemical compounds are not given in the work. However, an increase in yield in the second year of biochar application is recorded quite often due to an increase in the cation exchange capacity of soils, the availability of nutrients, the ability to retain water and a decrease in the mobility and toxicity of aluminum and manganese [296,298].
The direct or indirect effect of biochar on the growth of the root system and phytotoxic compounds indirectly affects crop yields. The acceleration of root culture development is due to the general improving effect of biochar on the soil structure, including its aeration and moisture saturation. Moreover, more than a hundred years ago, it was already believed that root development was accelerated due to the sorption removal of allelopathic compounds by activated carbons and similar products. At the same time, it should be taken into account that, most likely, the sorption removal of allelopaths and other phytotoxic compounds, including those released from litter and other plant residues, is complicated by the complex effect of biochar on the availability of nutrients, which leads to contradictory results in laboratory experiments [299]. In a number of studies, biochar was used as a substitute for peat in a nutrient medium for growing plants (soilless method). At a content of less than 30 wt%, biochar had a positive effect on plants, which was manifested in improved seed germination and seedling survival. In a number of studies, biochar was used as a substitute for peat in a substrate for growing plants (soilless method). Peat moss is conventionally used as a container substrate; however, this use causes several environmental concerns, such as peatland ecosystem destruction when peat is obtained, over-use of fertilizers since peat itself does not provide plants with essential nutrients, and carbon release since peat organic matter is quickly decomposed by plant-associated microbes. Biochar seems to have the potential to alleviate these concerns; however, the number of publications on peat substitution by biochar is still quite low as compared with the total number of publications on biochar used as a soil conditioner, suggesting that this study direction will be developed in the future. In most cases (depending on the initial substrate and pyrolysis process characteristics), biochar has a larger surface area, better pore size distribution, better rewetting characteristics, and higher content of N, P and other essential elements [300]. Examples of assessing the impact of biochar in the soilless method of cultivating a number of crops are shown in Table 10.
It is necessary to discuss separately the effect of biochar on plant diseases [308]. Even 170 years ago, there was a decrease in rust damage in wheat and powdery mildew in other crops. Since then, the effect of biochar has been tested on 13 pathogens, of which 85% have responded to biochar application by reducing plant infection, 12% have not, and 3% have increased plant disease. Later [309], the list of potential pathogens was expanded to 15 (30 plant/pathogen pathosystems). In 60% of the cases (70% of pathosystems), the maximum biochar concentration of 0.5–1% did not have any effect or had a negative effect on plant disease. Moreover, increasing the dose of biochar to 3% did not affect or accelerate the course of the disease. The study covered 12 different environments and 5 different prepared biochars, differing in salinity, carbon content, alkalinity and other parameters. The negative effect of biochar was less pronounced for foliar infections compared to soil pathogens. Other examples of assessing the impact of biochar on pathogens and plant infections are given in the review [310]. Examples of the acceleration of pathogen effects on plants are given in Table 11.
Strengthening the influence of pathogens under the effect of biochar often coincides with its suppressive effect on the development of the plants themselves, which manifests itself at high application rates. Moreover, even in the case of a positive effect on the growth rate of plants, the risk of irreversible damage to plants in the event of infection is potentially high. The contrary is also true—infected plants are more sensitive than healthy ones to an excess of applied biochar. The reasons for this phenomenon are not fully understood, but similar dose-dependent effects were observed for other agrochemicals, for example, glyphosate, which acts as a growth regulator at low concentrations, and as a herbicide at high concentrations. Biochar may contain certain organic components that, individually or in combination, have hormone-like or phytotoxic effects. Ethylene released from some biochars can serve as an example of such an indirect effect. In small doses, it has an effect on plants as a growth promoter and increases resistance to diseases, and in large doses, it has an opposite effect.

3.7. Immobilization of Microorganisms on Biochar

The number of positive examples of the use of biochar with immobilized microorganisms in soil bioremediation is rather limited. Thus, accelerated decomposition of the pyrethroid cypermethrin in soils enriched with biochar with an immobilized consortium of Bacillus zhanjiangensis (TJTB48), Bacillus pseudofirmus (TJTB58), and Oceanobacillus kimchii (TJTB66) bacteria was described [317]. Bacteria were grown on a mineral substrate and then mixed in an aqueous suspension with biochar (concentrations of 0.5, 1 and 2%) for 24 h. The mixture was added to the surface layer of soil contaminated with pyrethroid. An increase in the rate of decomposition of cypermethrin was observed at a biochar load of 0.5% (time to reduce the concentration by 2 times 16.4 days).
Accelerated bioremediation of soils contaminated with PAHs under the action of bacteria immobilized on a biochar was described [318]. To do this, we used bacteria previously isolated by the ability to degrade polyaromatic hydrocarbons (Pseudomonas putida and the second one remains unidentified). The microorganisms were immobilized on pre-crushed biomass (wood chips, bamboo leaves, orange bark, pine needles) and its pyrolysis product at 100, 300, 400 and 700 °C under oxygen-deficient conditions. Immobilization was carried out by mixing the cell suspension with the carrier in the presence of sodium alginate. Next, the mixture was dried and subjected to gelation by the introduction of calcium chloride. The presence of bacteria increased the degree of sorption extraction of hydrocarbons from water, while the organic mass before pyrolysis insignificantly reduced the bioavailability of the absorbed toxicant. When microorganisms immobilized on an organic substrate and biochar were placed in soils in a 90-day experiment, the highest rate of decomposition of polyaromatic hydrocarbons was demonstrated by a sample in which the bacteria carrier was biochar prepared at 400 °C.
Corynebacterium variabile HRJ4, which is highly salt tolerant, was used to accelerate the oxidation of gasoline hydrocarbons. The bacterium was immobilized on biochar prepared from wood chips by pyrolysis at 250, 400, or 700 °C [319]. The pyrolysis product was mixed with a suspension of microorganisms in a ratio of 5:100 for 12 h and then stored at 4 °C. The immobilized bacteria showed the highest activity in the oxidation of aliphatic hydrocarbons C16, C18, C19, C26 and C26, as well as pyrene and naphthalene. Activity during 7-day incubation reached 79% of the relatively free non-immobilized bacteria. Biochar from corn cobs and pig manure was used as a carrier of the mutant strain B. subtilis B38 in the study of bioremediation of soils contaminated with heavy metals [320]. It has been established that the presence of microorganisms contributes to a decrease in the bioavailability of cationic and anionic forms of metals (cadmium, lead, chromium) and increases the biomass during the cultivation of lettuce three times more than in the absence of microorganisms. The introduction of phosphate-mobilizing bacteria into biochar was used to reduce the toxicity of lead in soils [242]. For this purpose, Pseudomonal chlororaphis in the culture liquid was mixed in a volume ratio of 1:10 with biochar prepared from cow dung by pyrolysis at 200, 300, 400 and 500 °C for 4 h. The effect of additives was assessed by the degree of lead extraction from contaminated soils.

4. Conclusions

It can be concluded that the interest of both the scientific community and enterprises in biochar is growing. The variety of substrates for the production of biochar and the scope of its application are increasing. The characteristics of biochar are very diverse; it is possible to achieve the optimal ones both through the selection of the optimal substrate for pyrolysis, pyrolysis conditions, and through pre- or post-treatment. The use of biochar as a soil fertilizer leads to changes in the physical, chemical and biological properties of soils. This increases bulk density, soil porosity, and wet aggregate stability, while clayey soils decrease tensile strength. Biochar increases the pH and soluble carbon of soils, immobilizes toxicants, and reduces the bioavailability of toxins to soil organisms. However, biochar can contain phytotoxic compounds, such as PAHs and heavy metals. The use of biochar as a soil fertilizer leads to an increase in the amount of macro- and microelements in the soil. The introduction of biochar into the soil leads to an increase in the number and diversity of the soil microbial community, with the greatest effect observed for the rhizobiome. All of these factors influence plant growth and crop quality and consequently, more input of rhizodeposits into the soil if the dose of biochar application was right chosen. In addition, the production of biochar is zero-emission technology; its use as a soil amendment leads to the sequestration of carbon in the soil and greater efficiency in the capture of carbon dioxide in plant biomass. We assume that further research will be aimed at optimizing the efficiency of the pyrolysis process and finding methods for increasing the sequestration potential of biochar.

Funding

This work was funded by the subsidy allocated to Kazan Federal University for the state assignment in the sphere of scientific activities, project № FZSM-2022-0003.

Institutional Review Board Statement

Not applicable.

Data Availability Statement

Acknowledgments

The authors of the article express their sincere gratitude to Yakov Kuzyakov for the idea of the review and valuable comments during the preparation of the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. IBI. Standardized Product Definition and Product Testing Guidelines for Biochar That Is Used in Soil; International Biochar Initiative: Bowdoinham, ME, USA, 2015; Available online: http://www.biochar-international.org/characterizationstandard (accessed on 7 August 2023).
  2. Xu, G.; Lv, Y.; Sun, J.; Shao, H.; Wei, L. Recent Advances in Biochar Applications in Agricultural Soils: Benefits and Environmental Implications. CLEAN—Soil Air Water 2012, 40, 1093–1098. [Google Scholar] [CrossRef]
  3. Warnock, D.D.; Lehmann, J.; Kuyper, T.W.; Rillig, M.C. Mycorrhizal Responses to Biochar in Soil—Oncepts and Mechanisms; Springer: Berlin/Heidelberg, Germany, 2007; Volume 300, pp. 9–20. [Google Scholar]
  4. An, N.; Zhang, L.; Liu, Y.; Shen, S.; Li, N.; Wu, Z.; Yang, J.; Han, W.; Han, X. Biochar application with reduced chemical fertilizers improves soil pore structure and rice productivity. Chemosphere 2022, 298, 134304. [Google Scholar]
  5. Huang, Y.; Wang, C.; Lin, C.; Zhang, Y.; Chen, X.; Tang, L.; Liu, C.; Chen, Q.; Onwuka, M.I.; Song, T. Methane and Nitrous Oxide Flux after Biochar Application in Subtropical Acidic Paddy Soils under Tobacco-Rice Rotation. Sci. Rep. 2019, 9, 17277. [Google Scholar] [CrossRef] [PubMed]
  6. Tang, Z.; Liu, X.; Li, G.; Liu, X. Mechanism of biochar on nitrification and denitrification to N2O emissions based on isotope characteristic values. Environ. Res. 2022, 212, 113219. [Google Scholar] [CrossRef] [PubMed]
  7. Bruun, E.W.; Hauggaard-Nielsen, H.; Ibrahim, N.; Egsgaard, H.; Ambus, P.; Jensen, P.A.; Dam-Johansen, K. Influence of fast pyrolysis temperature on biochar labile fraction and short-term carbon loss in a loamy soil. Biomass. Bioenergy 2011, 35, 1182–1189. [Google Scholar] [CrossRef]
  8. Verheijen, F.G.A.; Graber, E.R.; Ameloot, N.; Bastos, A.C.; Sohi, S.; Knicker, H. Biochars in soils: New insights and emerging research needs. Eur. J. Soil Sci. 2014, 65, 22–27. [Google Scholar] [CrossRef]
  9. Awad, Y.M.; Blagodatskaya, E.; Ok, Y.S.; Kuzyakov, Y. Effects of polyacrylamide, biopolymer and biochar on the decomposition of 14C-labelled maize residues and on their stabilization in soil aggregates. Eur. J. Soil Biol. 2013, 64, 488–499. [Google Scholar] [CrossRef]
  10. Yin, J.; Zhao, L.; Xu, X.; Li, D.; Qiu, H.; Cao, X. Evaluation of long-term carbon sequestration of biochar in soil with biogeochemical field model. Sci. Total Environ. 2022, 822, 153576. [Google Scholar] [CrossRef]
  11. Kuryntseva, P.; Galitskaya, P.; Selivanovskaya, S. Optimization of pyrolysis regime for chicken manure treatment and biochar production. Water Environ. J. 2022, 36, 270–281. [Google Scholar] [CrossRef]
  12. Ennis, C.J.; Evans, A.G.; Islam, M.; Ralebitso-Senior, T.K.; Senior, E. Biochar: Carbon Sequestration, Land Remediation, and Impacts on Soil Icrobiology. Environ. Sustain. 2012, 42, 2311–2364. [Google Scholar] [CrossRef]
  13. Clough, T.J.; Condron, L.M. Biochar and the Nitrogen Cycle: Introduction. J. Environ. Qual. 2010, 39, 1218–1223. [Google Scholar] [CrossRef]
  14. Junna, S.; Bingchen, W.; Gang, X.; Hongbo, S. Effects of wheat straw biochar on carbon mineralization and guidance for large-scale soil quality improvement in the coastal wetland. Ecol. Eng. 2014, 62, 43–47. [Google Scholar] [CrossRef]
  15. Barrow, C.J. Biochar: Potential for countering land degradation and for improving agriculture. Appl. Geogr. 2012, 34, 21–28. [Google Scholar] [CrossRef]
  16. Haris, M.; Hamid, Y.; Usman, M.; Wang, L.; Saleem, A.; Su, F.; Guo, J.; Li, Y. Crop-residues derived biochar: Synthesis, properties, characterization and application for the removal of trace elements in soils. J. Hazard. Mater. 2021, 416, 126212. [Google Scholar] [CrossRef]
  17. Gao, J.; Zhao, Y.; Zhang, W.; Sui, Y.; Jin, D.; Xin, W.; Yi, J.; He, D. Biochar prepared at different pyrolysis temperatures affects urea-nitrogen immobilization and N2O emissions in paddy fields. PeerJ 2019, 2019, e7027. [Google Scholar] [CrossRef] [PubMed]
  18. Yu, K.L.; Lau, B.F.; Show, P.L.; Ong, H.C.; Ling, T.C.; Chen, W.H.; Ng, E.P.; Chang, J.S. Recent developments on algal biochar production and characterization. Bioresour. Technol. 2017, 246, 2–11. [Google Scholar] [CrossRef] [PubMed]
  19. Ronsse, F.; van Hecke, S.; Dickinson, D.; Prins, W. Production and characterization of slow pyrolysis biochar: Influence of feedstock type and pyrolysis conditions. GCB Bioenergy 2013, 5, 104–115. [Google Scholar] [CrossRef]
  20. Qi, L.; Pokharel, P.; Chang, S.X.; Zhou, P.; Niu, H.; He, X.; Wang, Z.; Gao, M. Biochar application increased methane emission, soil carbon storage and net ecosystem carbon budget in a 2-year vegetable–rice rotation. Agric. Ecosyst. Environ. 2020, 292, 106831. [Google Scholar] [CrossRef]
  21. Tsai, W.T.; Liu, S.C.; Chen, H.R.; Chang, Y.M.; Tsai, Y.L. Textural and chemical properties of swine-manure-derived biochar pertinent to its potential use as a soil amendment. Chemosphere 2012, 89, 198–203. [Google Scholar] [CrossRef]
  22. Cely, P.; Gascó, G.; Paz-Ferreiro, J.; Méndez, A. Agronomic properties of biochars from different manure wastes. J. Anal. Appl. Pyrolysis 2015, 111, 173–182. [Google Scholar] [CrossRef]
  23. Liu, W.J.; Jiang, H.; Yu, H.Q. Development of Biochar-Based Functional Materials: Toward a Sustainable Platform Carbon Material. Chem. Rev. 2015, 115, 12251–12285. [Google Scholar] [CrossRef]
  24. Xue, Y.; Wang, C.; Hu, Z.; Zhou, Y.; Xiao, Y.; Wang, T. Pyrolysis of sewage sludge by electromagnetic induction: Biochar properties and application in adsorption removal of Pb(II), Cd(II) from aqueous solution. Waste Manag. 2019, 89, 48–56. [Google Scholar] [CrossRef] [PubMed]
  25. Xu, G.; Zhang, Z.; Deng, L. Adsorption behaviors and removal efficiencies of inorganic, polymeric and organic phosphates from aqueous solution on biochar derived from sewage sludge of chemically enhanced primary treatment process. Water 2018, 10, 869. [Google Scholar] [CrossRef]
  26. Paz-Ferreiro, J.; Nieto, A.; Méndez, A.; Askeland, M.P.J.; Gascó, G. Biochar from biosolids pyrolysis: A review. Int. J. Environ. Res. Public Health 2018, 15, 956. [Google Scholar] [CrossRef]
  27. Vassilev, S.V.; Baxter, D.; Andersen, L.K.; Vassileva, C.G. An overview of the chemical composition of biomass. Fuel 2010, 89, 913–933. [Google Scholar] [CrossRef]
  28. Dhyani, V.; Bhaskar, T. A comprehensive review on the pyrolysis of lignocellulosic biomass. Renew. Energy 2018, 129, 695–716. [Google Scholar] [CrossRef]
  29. Krylova, A.Y.; Zaitchenko, V.M. Hydrothermal Carbonization of Biomass: A Review. Solid Fuel Chem. 2018, 52, 91–103. [Google Scholar] [CrossRef]
  30. Malherbe, S.; Cloete, T.E. Lignocellulose biodegradation: Fundamentals and applications. Rev. Environ. Sci. Bio/Technol. 2002, 1, 105–114. [Google Scholar] [CrossRef]
  31. Vassilev, S.V.; Baxter, D.; Andersen, L.K.; Vassileva, C.G.; Morgan, T.J. An overview of the organic and inorganic phase composition of biomass. Fuel 2012, 94, 1–33. [Google Scholar] [CrossRef]
  32. Demirbaş, A. Calculation of higher heating values of biomass fuels. Fuel 1997, 76, 431–434. [Google Scholar] [CrossRef]
  33. Raveendran, K.; Ganesh, A.; Khilar, K.C. Influence of mineral matter on biomass pyrolysis characteristics. Fuel 1995, 74, 1812–1822. [Google Scholar] [CrossRef]
  34. Saura-Calixto, F.; Cañellas, J.; Garcia-Raso, J. Determination of hemicellulose, cellulose and lignin contents of dietary fibre and crude fibre of several seed hulls. Data comparison. Z. Lebensm. Unters. Forsch. 1983, 177, 200–202. [Google Scholar] [CrossRef]
  35. Ververis, C.; Georghiou, K.; Danielidis, D.; Hatzinikolaou, D.G.; Santas, P.; Santas, R.; Corleti, V. Cellulose, hemicelluloses, lignin and ash content of some organic materials and their suitability for use as paper pulp supplements. Bioresour. Technol. 2007, 98, 296–301. [Google Scholar] [CrossRef] [PubMed]
  36. Ayeni, A.O.; Adeeyo, O.A.; Oresegun, O.M.; Oladimeji, T.E. Compositional analysis of lignocellulosic materials: Evaluation of an economically viable method suitable for woody and non-woody biomass. Am. J. Eng. Res. 2015, 4, 14–19. [Google Scholar]
  37. Anwar, Z.; Gulfraz, M.; Irshad, M. Agro-industrial lignocellulosic biomass a key to unlock the future bio-energy: A brief review. J. Radiat. Res. Appl. Sci. 2014, 7, 163–173. [Google Scholar] [CrossRef]
  38. Guimarães, J.L.; Frollini, E.; da Silva, C.G.; Wypych, F.; Satyanarayana, K.G. Characterization of banana, sugarcane bagasse and sponge gourd fibers of Brazil. Ind. Crops Prod. 2009, 30, 407–415. [Google Scholar] [CrossRef]
  39. Howard, R.L.; Abotsi, E.; Van Rensburg, E.L.J.; Howard, S. Lignocellulose biotechnology: Issues of bioconversion and enzyme production. Afr. J. Biotechnol. 2003, 2, 602–619. [Google Scholar] [CrossRef]
  40. Prasad, S.; Singh, A.; Joshi, H.C. Ethanol as an alternative fuel from agricultural, industrial and urban residues. Resour. Conserv. Recycl. 2007, 50, 1–39. [Google Scholar] [CrossRef]
  41. Kambo, H.S.; Dutta, A. A comparative review of biochar and hydrochar in terms of production, physico-chemical properties and applications. Renew. Sustain. Energy Rev. 2015, 45, 359–378. [Google Scholar] [CrossRef]
  42. Mani, S.; Sokhansanj, S.; Bi, X.; Turhollow, A. Economics of producing fuel pellets from biomass. Appl. Eng. Agric. 2006, 22, 421–426. [Google Scholar] [CrossRef]
  43. Roy, P.; Dias, G. Prospects for pyrolysis technologies in the bioenergy sector: A review. Renew. Sustain. Energy Rev. 2017, 77, 59–69. [Google Scholar] [CrossRef]
  44. Posom, J.; Saechua, W.; Sirisomboon, P. Evaluation of pyrolysis characteristics of milled bamboo using near-infrared spectroscopy. Renew. Energy 2017, 103, 653–665. [Google Scholar] [CrossRef]
  45. Blanco-Canqui, H. Crop Residue Removal for Bioenergy Reduces Soil Carbon Pools: How Can We Offset Carbon Losses? Bioenergy Res. 2013, 6, 358–371. [Google Scholar] [CrossRef]
  46. Fuertes, A.B.; Arbestain, M.C.; Sevilla, M.; MacIá-Agulló, J.A.; Fiol, S.; López, R.; Smernik, R.J.; Aitkenhead, W.P.; Arce, F.; MacIas, F. Chemical and structural properties of carbonaceous products obtained by pyrolysis and hydrothermal carbonisation of corn stover. Aust. J. Soil Res. 2010, 48, 618–626. [Google Scholar] [CrossRef]
  47. Zhao, B.; Wang, X.; Yang, X. Co-pyrolysis characteristics of microalgae Isochrysis and Chlorella: Kinetics, biocrude yield and interaction. Bioresour. Technol. 2015, 198, 332–339. [Google Scholar] [CrossRef]
  48. Norouzi, O.; Jafarian, S.; Safari, F.; Tavasoli, A.; Nejati, B. Promotion of hydrogen-rich gas and phenolic-rich bio-oil production from green macroalgae Cladophora glomerata via pyrolysis over its bio-char. Bioresour. Technol. 2016, 219, 643–651. [Google Scholar] [CrossRef] [PubMed]
  49. Ma, W.; Du, G.; Li, J.; Fang, Y.; Hou, L.; Chen, G.; Ma, D. Supercritical water pyrolysis of sewage sludge. Waste Manag. 2017, 59, 371–378. [Google Scholar] [CrossRef]
  50. Calisto, V.; Ferreira, C.I.A.; Santos, S.M.; Gil, M.V.; Otero, M.; Esteves, V.I. Production of adsorbents by pyrolysis of paper mill sludge and application on the removal of citalopram from water. Bioresour. Technol. 2014, 166, 335–344. [Google Scholar] [CrossRef] [PubMed]
  51. Heilmann, S.M.; Molde, J.S.; Timler, J.G.; Wood, B.M.; Mikula, A.L.; Vozhdayev, G.V.; Colosky, E.C.; Spokas, K.A.; Valentas, K.J. Phosphorus reclamation through hydrothermal carbonization of animal manures. Environ. Sci. Technol. 2014, 48, 10323–10329. [Google Scholar] [CrossRef] [PubMed]
  52. Wu, W.; Yang, M.; Feng, Q.; McGrouther, K.; Wang, H.; Lu, H.; Chen, Y. Chemical characterization of rice straw-derived biochar for soil amendment. Biomass Bioenergy 2012, 47, 268–276. [Google Scholar] [CrossRef]
  53. Yuan, J.-H.H.; Xu, R.-K.K. Effects of biochars generated from crop residues on chemical properties of acid soils from tropical and subtropical China. Soil Res. 2012, 50, 570. [Google Scholar] [CrossRef]
  54. Imam, T.; Capareda, S. Characterization of bio-oil, syn-gas and bio-char from switchgrass pyrolysis at various temperatures. J. Anal. Appl. Pyrolysis 2012, 93, 170–177. [Google Scholar] [CrossRef]
  55. Kinney, T.J.; Masiello, C.A.; Dugan, B.; Hockaday, W.C.; Dean, M.R.; Zygourakis, K.; Barnes, R.T. Hydrological properties of biochars produced at different temperatures. Biomass Bioenergy 2012, 41, 34–43. [Google Scholar] [CrossRef]
  56. Uzoma, K.C.; Inoue, M.; Andry, H.; Fujimaki, H.; Zahoor, A.; Nishihara, E. Effect of cow manure biochar on maize productivity under sandy soil condition. Soil Use Manag. 2011, 27, 205–212. [Google Scholar] [CrossRef]
  57. Mašek, O.; Budarin, V.; Gronnow, M.; Crombie, K.; Brownsort, P.; Fitzpatrick, E.; Hurst, P. Microwave and slow pyrolysis biochar—Comparison of physical and functional properties. J. Anal. Appl. Pyrolysis 2013, 100, 41–48. [Google Scholar] [CrossRef]
  58. Zhao, L.; Cao, X.; Mašek, O.; Zimmerman, A. Heterogeneity of biochar properties as a function of feedstock sources and production temperatures. J. Hazard. Mater. 2013, 256–257, 1–9. [Google Scholar] [CrossRef]
  59. Kim, K.H.; Eom, I.Y.; Lee, S.M.; Choi, D.; Yeo, H.; Choi, I.-G.; Choi, J.W. Investigation of physicochemical properties of biooils produced from yellow poplar wood (Liriodendron tulipifera) at various temperatures and residence times. J. Anal. Appl. Pyrolysis 2011, 92, 2–9. [Google Scholar] [CrossRef]
  60. Laghari, M.; Hu, Z.; Mirjat, M.S.; Xiao, B.; Tagar, A.A.; Hu, M. Fast pyrolysis biochar from sawdust improves the quality of desert soils and enhances plant growth. J. Sci. Food Agric. 2016, 96, 199–206. [Google Scholar] [CrossRef] [PubMed]
  61. Song, W.; Guo, M. Quality variations of poultry litter biochar generated at different pyrolysis temperatures. J. Anal. Appl. Pyrolysis 2012, 94, 138–145. [Google Scholar] [CrossRef]
  62. Godwin, P.M.; Pan, Y.; Xiao, H.; Afzal, M.T. Progress in Preparation and Application of Modified Biochar for Improving Heavy Metal Ion Removal From Wastewater. J. Bioresour. Bioprod. 2019, 4, 31–42. [Google Scholar] [CrossRef]
  63. McKendry, P. Energy production from biomass (part 1): Overview of biomass. Bioresour. Technol. 2002, 83, 37–46. [Google Scholar] [CrossRef] [PubMed]
  64. Mumme, J.; Eckervogt, L.; Pielert, J.; Diakité, M.; Rupp, F.; Kern, J. Hydrothermal carbonization of anaerobically digested maize silage. Bioresour. Technol. 2011, 102, 9255–9260. [Google Scholar] [CrossRef] [PubMed]
  65. Hoekman, S.K.; Broch, A.; Felix, L.; Farthing, W. Hydrothermal carbonization (HTC) of loblolly pine using a continuous, reactive twin-screw extruder. Energy Convers. Manag. 2017, 134, 247–259. [Google Scholar] [CrossRef]
  66. Ho, S.H.; Zhang, C.; Chen, W.H.; Shen, Y.; Chang, J.S. Characterization of biomass waste torrefaction under conventional and microwave heating. Bioresour. Technol. 2018, 264, 7–16. [Google Scholar] [CrossRef] [PubMed]
  67. Weber, K.; Quicker, P. Properties of Biochar; Elsevier Ltd.: Amsterdam, The Netherlands, 2018; Volume 217, pp. 240–261. [Google Scholar]
  68. Bruun, E.W.; Ambus, P.; Egsgaard, H.; Hauggaard-Nielsen, H. Effects of slow and fast pyrolysis biochar on soil C and N turnover dynamics. Soil Biol. Biochem. 2012, 46, 73–79. [Google Scholar] [CrossRef]
  69. Zhao, L.; Cao, X.; Wang, Q.; Yang, F.; Xu, S. Mineral Constituents Profile of Biochar Derived from Diversified Waste Biomasses: Implications for Agricultural Applications. J. Environ. Qual. 2013, 42, 545–552. [Google Scholar] [CrossRef]
  70. Nunoura, T.; Wade, S.R.; Bourke, J.P.; Antal, M.J. Studies of the flash carbonization process. 1. Propagation of the flaming pyrolysis reaction and performance of a catalytic afterburner. Ind. Eng. Chem. Res. 2006, 45, 585–599. [Google Scholar] [CrossRef]
  71. Bourke, J.; Manley-Harris, M.; Fushimi, C.; Dowaki, K.; Nunoura, T.; Antal, M.J. Do all carbonized charcoals have the same chemical structure? 2. A model of the chemical structure of carbonized charcoal. Ind. Eng. Chem. Res. 2007, 46, 5954–5967. [Google Scholar] [CrossRef]
  72. Sabio, E.; Álvarez-Murillo, A.; Román, S.; Ledesma, B. Conversion of tomato-peel waste into solid fuel by hydrothermal carbonization: Influence of the processing variables. Waste Manag. 2016, 47, 122–132. [Google Scholar] [CrossRef]
  73. Lynam, J.G.; Coronella, C.J.; Yan, W.; Reza, M.T.; Vasquez, V.R. Acetic acid and lithium chloride effects on hydrothermal carbonization of lignocellulosic biomass. Bioresour. Technol. 2011, 102, 6192–6199. [Google Scholar] [CrossRef]
  74. Lynam, J.G.; Toufiq Reza, M.; Vasquez, V.R.; Coronella, C.J. Effect of salt addition on hydrothermal carbonization of lignocellulosic biomass. Fuel 2012, 99, 271–273. [Google Scholar] [CrossRef]
  75. Babu, B.V. Biomass pyrolysis: A state-of-the-art review. Biofuels Bioprod. Biorefining 2008, 2, 393–414. [Google Scholar] [CrossRef]
  76. Shen, D.K.; Gu, S. The mechanism for thermal decomposition of cellulose and its main products. Bioresour. Technol. 2009, 100, 6496–6504. [Google Scholar] [CrossRef] [PubMed]
  77. Patwardhan, P.R.; Brown, R.C.; Shanks, B.H. Product distribution from the fast pyrolysis of hemicellulose. ChemSusChem 2011, 4, 636–643. [Google Scholar] [CrossRef]
  78. Mu, W.; Ben, H.; Ragauskas, A.; Deng, Y. Lignin Pyrolysis Components and Upgrading-Technology Review. Bioenergy Res. 2013, 6, 1183–1204. [Google Scholar] [CrossRef]
  79. Xin-hui, D.; Srinivasakannan, C.; Jin-hui, P.; Li-bo, Z.; Zheng-yong, Z. Comparison of activated carbon prepared from Jatropha hull by conventional heating and microwave heating. Biomass Bioenergy 2011, 35, 3920–3926. [Google Scholar] [CrossRef]
  80. Jagtoyen, M.; Derbyshire, F. Activated carbons from yellow poplar and white oak by H3PO4 activation. Carbon 1998, 36, 1085–1097. [Google Scholar] [CrossRef]
  81. Zhao, L.; Zheng, W.; Mašek, O.; Chen, X.; Gu, B.; Sharma, B.K.; Cao, X. Roles of Phosphoric Acid in Biochar Formation: Synchronously Improving Carbon Retention and Sorption Capacity. J. Environ. Qual. 2017, 46, 393–401. [Google Scholar] [CrossRef] [PubMed]
  82. Uçar, S.; Erdem, M.; Tay, T.; Karagöz, S. Preparation and characterization of activated carbon produced from pomegranate seeds by ZnCl2 activation. Appl. Surf. Sci. 2009, 255, 8890–8896. [Google Scholar] [CrossRef]
  83. Yaman, S. Pyrolysis of biomass to produce fuels and chemical feedstocks. Energy Convers. Manag. 2004, 45, 651–671. [Google Scholar] [CrossRef]
  84. Liu, N.; Zhou, J.; Han, L.; Ma, S.; Sun, X.; Huang, G. Role and multi-scale characterization of bamboo biochar during poultry manure aerobic composting. Bioresour. Technol. 2017, 241, 190–199. [Google Scholar] [CrossRef] [PubMed]
  85. Jansen, R.J.J.; van Bekkum, H. Amination and ammoxidation of activated carbons. Carbon 1994, 32, 1507–1516. [Google Scholar] [CrossRef]
  86. Shafeeyan, M.S.; Wan Daud, W.M.A.; Houshmand, A.; Arami-Niya, A. The application of response surface methodology to optimize the amination of activated carbon for the preparation of carbon dioxide adsorbents. Fuel 2012, 94, 465–472. [Google Scholar] [CrossRef]
  87. Anfruns, A.; García-Suárez, E.J.; Montes-Morán, M.A.; Gonzalez-Olmos, R.; Martin, M.J. New insights into the influence of activated carbon surface oxygen groups on H2O2 decomposition and oxidation of pre-adsorbed volatile organic compounds. Carbon 2014, 77, 89–98. [Google Scholar] [CrossRef]
  88. Wu, X.; Xu, J.; Dong, F.; Liu, X.; Zheng, Y. Responses of soil microbial community to different concentration of fomesafen. J. Hazard. Mater. 2014, 273, 155–164. [Google Scholar] [CrossRef]
  89. Gokce, Y.; Aktas, Z. Nitric acid modification of activated carbon produced from waste tea and adsorption of methylene blue and phenol. Appl. Surf. Sci. 2014, 313, 352–359. [Google Scholar] [CrossRef]
  90. Tu, C.; Wei, J.; Guan, F.; Liu, Y.; Sun, Y.; Luo, Y. Biochar and bacteria inoculated biochar enhanced Cd and Cu immobilization and enzymatic activity in a polluted soil. Environ. Int. 2020, 137, 105576. [Google Scholar] [CrossRef]
  91. Jung, K.W.; Kim, K.; Jeong, T.U.; Ahn, K.H. Influence of pyrolysis temperature on characteristics and phosphate adsorption capability of biochar derived from waste-marine macroalgae (Undaria pinnatifida roots). Bioresour. Technol. 2016, 200, 1024–1028. [Google Scholar] [CrossRef]
  92. Fletcher, A.J.; Smith, M.A.; Heinemeyer, A.; Lord, R.; Ennis, C.J.; Hodgson, E.M.; Farrar, K. Production Factors Controlling the Physical Characteristics of Biochar Derived from Phytoremediation Willow for Agricultural Applications. Bioenergy Res. 2014, 7, 371–380. [Google Scholar] [CrossRef]
  93. Dehkhoda, A.M.; West, A.H.; Ellis, N. Biochar based solid acid catalyst for biodiesel production. Appl. Catal. A Gen. 2010, 382, 197–204. [Google Scholar] [CrossRef]
  94. Yang, K.; Peng, J.; Srinivasakannan, C.; Zhang, L.; Xia, H.; Duan, X. Preparation of high surface area activated carbon from coconut shells using microwave heating. Bioresour. Technol. 2010, 101, 6163–6169. [Google Scholar] [CrossRef]
  95. Zhao, S.X.; Ta, N.; Wang, X.D. Effect of temperature on the structural and physicochemical properties of biochar with apple tree branches as feedstock material. Energies 2017, 10, 1293. [Google Scholar] [CrossRef]
  96. Rajapaksha, A.U.; Chen, S.S.; Tsang, D.C.W.; Zhang, M.; Vithanage, M.; Mandal, S.; Gao, B.; Bolan, N.S.; Ok, Y.S. Engineered/designer biochar for contaminant removal/immobilization from soil and water: Potential and implication of biochar modification. Chemosphere 2016, 148, 276–291. [Google Scholar] [CrossRef]
  97. Frišták, V.; Micháleková-Richveisová, B.; Víglašová, E.; Ďuriška, L.; Galamboš, M.; Moreno-Jimenéz, E.; Pipíška, M.; Soja, G. Sorption separation of Eu and As from single-component systems by Fe-modified biochar: Kinetic and equilibrium study. J. Iran. Chem. Soc. 2017, 14, 521–530. [Google Scholar] [CrossRef]
  98. Wang, S.; Gao, B.; Zimmerman, A.R.; Li, Y.; Ma, L.; Harris, W.G.; Migliaccio, K.W. Removal of arsenic by magnetic biochar prepared from pinewood and natural hematite. Bioresour. Technol. 2015, 175, 391–395. [Google Scholar] [CrossRef] [PubMed]
  99. Zhou, Z.; Liu, Y.G.; Liu, S.B.; Liu, H.Y.; Zeng, G.M.; Tan, X.F.; Yang, C.P.; Ding, Y.; Yan, Z.L.; Cai, X.X. Sorption performance and mechanisms of arsenic(V) removal by magnetic gelatin-modified biochar. Chem. Eng. J. 2017, 314, 223–231. [Google Scholar] [CrossRef]
  100. Zhang, M.; Gao, B. Removal of arsenic, methylene blue, and phosphate by biochar/AlOOH nanocomposite. Chem. Eng. J. 2013, 226, 286–292. [Google Scholar] [CrossRef]
  101. Wang, S.; Gao, B.; Li, Y.; Mosa, A.; Zimmerman, A.R.; Ma, L.Q.; Harris, W.G.; Migliaccio, K.W. Manganese oxide-modified biochars: Preparation, characterization, and sorption of arsenate and lead. Bioresour. Technol. 2015, 181, 13–17. [Google Scholar] [CrossRef] [PubMed]
  102. Zhang, M.; Gao, B.; Yao, Y.; Xue, Y.; Inyang, M. Synthesis of porous MgO-biochar nanocomposites for removal of phosphate and nitrate from aqueous solutions. Chem. Eng. J. 2012, 210, 26–32. [Google Scholar] [CrossRef]
  103. Yao, Y.; Gao, B.; Chen, J.; Zhang, M.; Inyang, M.; Li, Y.; Alva, A.; Yang, L. Engineered carbon (biochar) prepared by direct pyrolysis of Mg-accumulated tomato tissues: Characterization and phosphate removal potential. Bioresour. Technol. 2013, 138, 8–13. [Google Scholar] [CrossRef]
  104. Ismadji, S.; Tong, D.S.; Soetaredjo, F.E.; Ayucitra, A.; Yu, W.H.; Zhou, C.H. Bentonite hydrochar composite for removal of ammonium from Koi fish tank. Appl. Clay Sci. 2016, 119, 146–154. [Google Scholar] [CrossRef]
  105. Tan, X.F.; Liu, Y.G.; Gu, Y.L.; Xu, Y.; Zeng, G.M.; Hu, X.J.; Liu, S.B.; Wang, X.; Liu, S.M.; Li, J. Biochar-based nano-composites for the decontamination of wastewater: A review. Bioresour. Technol. 2016, 212, 318–333. [Google Scholar] [CrossRef]
  106. Shang, M.R.; Liu, Y.G.; Liu, S.B.; Zeng, G.M.; Tan, X.F.; Jiang, L.H.; Huang, X.X.; Ding, Y.; Guo, Y.M.; Wang, S.F. A novel graphene oxide coated biochar composite: Synthesis, characterization and application for Cr(VI) removal. RSC Adv. 2016, 6, 85202–85212. [Google Scholar] [CrossRef]
  107. Inyang, M.; Gao, B.; Zimmerman, A.; Zhou, Y.; Cao, X. Sorption and cosorption of lead and sulfapyridine on carbon nanotube-modified biochars. Environ. Sci. Pollut. Res. 2015, 22, 1868–1876. [Google Scholar] [CrossRef]
  108. Downie, A.; Crosky, A.; Munroe, P. Physical properties of biochar. In Biochar for Environmental Management: Science and Technology; Routledge: Abingdon, UK, 2012; ISBN 9781849770552. [Google Scholar]
  109. Shen, D.K.; Gu, S.; Bridgwater, A.V. Study on the pyrolytic behaviour of xylan-based hemicellulose using TG-FTIR and Py-GC-FTIR. J. Anal. Appl. Pyrolysis 2010, 87, 199–206. [Google Scholar] [CrossRef]
  110. Kappler, A.; Wuestner, M.L.; Ruecker, A.; Harter, J.; Halama, M.; Behrens, S. Biochar as an Electron Shuttle between Bacteria and Fe(III) Minerals. Environ. Sci. Technol. Lett. 2014, 1, 339–344. [Google Scholar] [CrossRef]
  111. Baccile, N.; Laurent, G.; Babonneau, F.; Fayon, F.; Titirici, M.M.; Antonietti, M. Structural characterization of hydrothermal carbon spheres by advanced solid-state MAS 13C NMR investigations. J. Phys. Chem. C 2009, 113, 9644–9654. [Google Scholar] [CrossRef]
  112. Pastorova, I.; Botto, R.E.; Arisz, P.W.; Boon, J.J. Cellulose char structure: A combined analytical Py-GC-MS, FTIR, and NMR study. Carbohydr. Res. 1994, 262, 27–47. [Google Scholar] [CrossRef]
  113. Peters, B. Prediction of pyrolysis of pistachio shells based on its components hemicellulose, cellulose and lignin. Fuel Process. Technol. 2011, 92, 1993–1998. [Google Scholar] [CrossRef]
  114. Huang, J.; Liu, C.; Tong, H.; Li, W.; Wu, D. Theoretical studies on pyrolysis mechanism of xylopyranose. Comput. Theor. Chem. 2012, 1001, 44–50. [Google Scholar] [CrossRef]
  115. Shafizadeh, F. Introduction to pyrolysis of biomass. J. Anal. Appl. Pyrolysis 1982, 3, 283–305. [Google Scholar] [CrossRef]
  116. Wornat, M.J.; Hurt, R.H.; Yang, N.Y.C.; Headley, T.J. Structural and compositional transformations of biomass chars during combustion. Combust. Flame 1995, 100, 131–143. [Google Scholar] [CrossRef]
  117. Brewer, C.E.; Schmidt-Rohr, K.; Satrio, J.A.; Brown, R.C. Characterization of biochar from fast pyrolysis and gasification systems. Environ. Prog. Sustain. Energy 2009, 28, 386–396. [Google Scholar] [CrossRef]
  118. Lua, A.C.; Yang, T.; Guo, J. Effects of pyrolysis conditions on the properties of activated carbons prepared from pistachio-nut shells. J. Anal. Appl. Pyrolysis 2004, 72, 279–287. [Google Scholar] [CrossRef]
  119. Plötze, M.; Niemz, P. Porosity and pore size distribution of different wood types as determined by mercury intrusion porosimetry. Eur. J. Wood Wood Prod. 2011, 69, 649–657. [Google Scholar] [CrossRef]
  120. Brewer, C.E.; Chuang, V.J.; Masiello, C.A.; Gonnermann, H.; Gao, X.; Dugan, B.; Driver, L.E.; Panzacchi, P.; Zygourakis, K.; Davies, C.A. New approaches to measuring biochar density and porosity. Biomass Bioenergy 2014, 66, 176–185. [Google Scholar] [CrossRef]
  121. Pulido-Novicio, L.; Hata, T.; Kurimoto, Y.; Doi, S.; Ishihara, S.; Imamura, Y. Adsorption capacities and related characteristics of wood charcoals carbonized using a one-step or two-step process. J. Wood Sci. 2001, 47, 48–57. [Google Scholar] [CrossRef]
  122. Jimenez-Cordero, D.; Heras, F.; Alonso-Morales, N.; Gilarranz, M.A.; Rodriguez, J.J. Porous structure and morphology of granular chars from flash and conventional pyrolysis of grape seeds. Biomass Bioenergy 2013, 54, 123–132. [Google Scholar] [CrossRef]
  123. Cetin, E.; Gupta, R.; Moghtaderi, B. Effect of pyrolysis pressure and heating rate on radiata pine char structure and apparent gasification reactivity. Fuel 2005, 84, 1328–1334. [Google Scholar] [CrossRef]
  124. Wafiq, A.; Reichel, D.; Hanafy, M. Pressure influence on pyrolysis product properties of raw and torrefied Miscanthus: Role of particle structure. Fuel 2016, 179, 156–167. [Google Scholar] [CrossRef]
  125. Fu, P.; Hu, S.; Xiang, J.; Sun, L.; Li, P.; Zhang, J.; Zheng, C. Pyrolysis of maize stalk on the characterization of chars formed under different devolatilization conditions. Energy Fuels 2009, 23, 4605–4611. [Google Scholar] [CrossRef]
  126. Burhenne, L.; Damiani, M.; Aicher, T. Effect of feedstock water content and pyrolysis temperature on the structure and reactivity of spruce wood char produced in fixed bed pyrolysis. Fuel 2013, 107, 836–847. [Google Scholar] [CrossRef]
  127. Chen, Y.; Yang, H.; Wang, X.; Zhang, S.; Chen, H. Biomass-based pyrolytic polygeneration system on cotton stalk pyrolysis: Influence of temperature. Bioresour. Technol. 2012, 107, 411–418. [Google Scholar] [CrossRef]
  128. Yuan, H.; Lu, T.; Huang, H.; Zhao, D.; Kobayashi, N.; Chen, Y. Influence of pyrolysis temperature on physical and chemical properties of biochar made from sewage sludge. J. Anal. Appl. Pyrolysis 2015, 112, 284–289. [Google Scholar] [CrossRef]
  129. Shaaban, A.; Se, S.-M.; Dimin, M.F.; Juoi, J.M.; Mohd Husin, M.H.; Mitan, N.M.M. Influence of heating temperature and holding time on biochars derived from rubber wood sawdust via slow pyrolysis. J. Anal. Appl. Pyrolysis 2014, 107, 31–39. [Google Scholar] [CrossRef]
  130. Keiluweit, M.; Nico, P.S.; Johnson, M.; Kleber, M. Dynamic molecular structure of plant biomass-derived black carbon (biochar). Environ. Sci. Technol. 2010, 44, 1247–1253. [Google Scholar] [CrossRef]
  131. He, P.; Liu, Y.; Shao, L.; Zhang, H.; Lü, F. Particle size dependence of the physicochemical properties of biochar. Chemosphere 2018, 212, 385–392. [Google Scholar] [CrossRef] [PubMed]
  132. Abdullah, H.; Wu, H. Biochar as a fuel: 1. Properties and grindability of biochars produced from the pyrolysis of mallee wood under slow-heating conditions. Energy Fuels 2009, 23, 4174–4181. [Google Scholar] [CrossRef]
  133. Kim, K.H.; Kim, J.Y.; Cho, T.S.; Choi, J.W. Influence of pyrolysis temperature on physicochemical properties of biochar obtained from the fast pyrolysis of pitch pine (Pinus rigida). Bioresour. Technol. 2012, 118, 158–162. [Google Scholar] [CrossRef]
  134. Van Krevelen, D. Graphical statistical method for the study of structure and reaction processes of coal. Fuel 1950, 29, 269–284. [Google Scholar]
  135. Weber, K.; Quicker, P. Eigenschaften von Biomassekarbonisaten. In Biokohle; Springer: Wiesbaden, Germany, 2016. [Google Scholar]
  136. Mullen, C.A.; Boateng, A.A.; Goldberg, N.M.; Lima, I.M.; Laird, D.A.; Hicks, K.B. Bio-oil and bio-char production from corn cobs and stover by fast pyrolysis. Biomass Bioenergy 2010, 34, 67–74. [Google Scholar] [CrossRef]
  137. Nguyen, B.T.; Lehmann, J.; Hockaday, W.C.; Joseph, S.; Masiello, C.a. Temperature Sensitivity of Black Carbon Descomposition and Oxidation. Environ. Sci. Technol. 2010, 44, 3324–3331. [Google Scholar] [CrossRef] [PubMed]
  138. Cheng, C.-H.H.; Lehmann, J.; Thies, J.E.; Burton, S.D.; Engelhard, M.H. Oxidation of black carbon by biotic and abiotic processes. Org. Geochem. 2006, 37, 1477–1488. [Google Scholar] [CrossRef]
  139. Prasad, M.; Chrysargyris, A.; McDaniel, N.; Kavanagh, A.; Gruda, N.S.; Tzortzakis, N. Plant nutrient availability and pH of biochars and their fractions, with the possible use as a component in a growing media. Agronomy 2020, 10, 10. [Google Scholar] [CrossRef]
  140. Choppala, G.K.; Bolan, N.S.; Megharaj, M.; Chen, Z.; Naidu, R. The influence of biochar and black carbon on reduction and bioavailability of chromate in soils. J. Environ. Qual. 2012, 41, 1175. [Google Scholar] [CrossRef] [PubMed]
  141. Usowicz, B.; Lipiec, J.; Łukowski, M.; Marczewski, W.; Usowicz, J. The effect of biochar application on thermal properties and albedo of loess soil under grassland and fallow. Soil Tillage Res. 2016, 164, 45–51. [Google Scholar] [CrossRef]
  142. Githinji, L. Effect of biochar application rate on soil physical and hydraulic properties of a sandy loam. Arch. Agron. Soil Sci. 2014, 60, 457–470. [Google Scholar] [CrossRef]
  143. Głąb, T.; Palmowska, J.; Zaleski, T.; Gondek, K. Effect of biochar application on soil hydrological properties and physical quality of sandy soil. Geoderma 2016, 281, 11–20. [Google Scholar] [CrossRef]
  144. Liu, Z.; Dugan, B.; Masiello, C.A.; Barnes, R.T.; Gallagher, M.E.; Gonnermann, H. Impacts of biochar concentration and particle size on hydraulic conductivity and DOC leaching of biochar-sand mixtures. J. Hydrol. 2016, 533, 461–472. [Google Scholar] [CrossRef]
  145. Ouyang, L.; Wang, F.; Tang, J.; Yu, L.; Zhang, R. Effects of biochar amendment on soil aggregates and hydraulic properties. J. Soil Sci. Plant Nutr. 2013, 13, 991–1002. [Google Scholar] [CrossRef]
  146. Burrell, L.D.; Zehetner, F.; Rampazzo, N.; Wimmer, B.; Soja, G. Long-term effects of biochar on soil physical properties. Geoderma 2016, 282, 96–102. [Google Scholar] [CrossRef]
  147. Curaqueo, G.; Meier, S.; Khan, N.; Cea, M.; Navia, R. Use of biochar on two volcanic soils: Effects on soil properties and barley yield. J. Soil Sci. Plant Nutr. 2014, 14, 911–924. [Google Scholar] [CrossRef]
  148. Herath, H.M.S.K.; Camps-Arbestain, M.; Hedley, M. Effect of biochar on soil physical properties in two contrasting soils: An Alfisol and an Andisol. Geoderma 2013, 209–210, 188–197. [Google Scholar] [CrossRef]
  149. Gamage, D.N.V.; Mapa, R.B.; Dharmakeerthi, R.S.; Biswas, A. Effect of rice-husk biochar on selected soil properties in tropical Alfisols. Soil Res. 2016, 54, 302. [Google Scholar] [CrossRef]
  150. Liu, X.H.; Han, F.P.; Zhang, X.C. Effect of biochar on soil aggregates in the Loess Plateau: Results from incubation experiments. Int. J. Agric. Biol. 2012, 14, 975–979. [Google Scholar]
  151. Blanco-Canqui, H.; Lal, R.; Post, W.M.; Izaurralde, R.C.; Shipitalo, M.J. Organic Carbon Influences on Soil Particle Density and Rheological Properties. Soil Sci. Soc. Am. J. 2006, 70, 1407–1414. [Google Scholar] [CrossRef]
  152. Haque, A.N.A.; Uddin, M.K.; Sulaiman, M.F.; Amin, A.M.; Hossain, M.; Aziz, A.A.; Mosharrof, M. Impact of organic amendment with alternate wetting and drying irrigation on rice yield, water use efficiency and physicochemical properties of soil. Agronomy 2021, 11, 1529. [Google Scholar] [CrossRef]
  153. Zilberman, D.; Laird, D.; Rainey, C.; Song, J.; Kahn, G. Biochar supply-chain and challenges to commercialization. GCB Bioenergy 2023, 15, 7–23. [Google Scholar] [CrossRef]
  154. Andrenelli, M.C.; Maienza, A.; Genesio, L.; Miglietta, F.; Pellegrini, S.; Vaccari, F.P.; Vignozzi, N. Field application of pelletized biochar: Short term effect on the hydrological properties of a silty clay loam soil. Agric. Water Manag. 2016, 163, 190–196. [Google Scholar] [CrossRef]
  155. Chan, K.Y.; Van Zwieten, L.; Meszaros, I.; Downie, A.; Joseph, S. Agronomic values of greenwaste biochar as a soil amendment. Soil Res. 2007, 45, 629. [Google Scholar] [CrossRef]
  156. Khademalrasoul, A.; Naveed, M.; Heckrath, G.; Kumari, K.G.I.D.; Wollesen De Jonge, L.; Elsgaard, L.; Vogel, H.J.; Iversen, B.V. Biochar effects on soil aggregate properties under no-till maize. Soil Sci. 2014, 179, 273–283. [Google Scholar] [CrossRef]
  157. Zhang, H.; Lin, K.; Wang, H.; Gan, J. Effect of Pinus radiata derived biochars on soil sorption and desorption of phenanthrene. Environ. Pollut. 2010, 158, 2821–2825. [Google Scholar] [CrossRef] [PubMed]
  158. Beesley, L.; Moreno-Jiménez, E.; Gomez-Eyles, J.L.; Harris, E.; Robinson, B.; Sizmur, T. A review of biochars’ potential role in the remediation, revegetation and restoration of contaminated soils. Environ. Pollut. 2011, 159, 3269–3282. [Google Scholar] [CrossRef] [PubMed]
  159. Mohan, D.; Sarswat, A.; Ok, Y.S.; Pittman, C.U. Organic and inorganic contaminants removal from water with biochar, a renewable, low cost and sustainable adsorbent-A critical review. Bioresour. Technol. 2014, 160, 191–202. [Google Scholar] [CrossRef] [PubMed]
  160. Jeffery, S.; Bezemer, T.M.; Cornelissen, G.; Kuyper, T.W.; Lehmann, J.; Mommer, L.; Sohi, S.P.; van de Voorde, T.F.J.; Wardle, D.A.; van Groenigen, J.W. The way forward in biochar research: Targeting trade-offs between the potential wins. GCB Bioenergy 2015, 7, 1–13. [Google Scholar] [CrossRef]
  161. Sun, X.; Ghosh, U. The effect of activated carbon on partitioning, desorption, and biouptake of native polychlorinated biphenyls in four freshwater sediments. Environ. Toxicol. Chem. 2008, 27, 2287–2295. [Google Scholar] [CrossRef]
  162. Cho, Y.-M.; Ghosh, U.; Kennedy, A.J.; Grossman, A.; Ray, G.; Tomaszewski, J.E.; Smithenry, D.W.; Bridges, T.S.; Luthy, R.G. Field application of activated carbon amendment for in-situ stabilization of polychlorinated biphenyls in marine sediment. Environ. Sci. Technol. 2009, 43, 3815–3823. [Google Scholar] [CrossRef]
  163. Tomaszewski, J.E.; Werner, D.; Luthy, R.G. Activated carbon amendment as a treatment for residual DDT in sediment from a superfund site in San Francisco Bay, Richmond, California, USA. Environ. Toxicol. Chem. 2007, 26, 2143. [Google Scholar] [CrossRef]
  164. Zimmerman, J.R.; Ghosh, U.; Millward, R.N.; Bridges, T.S.; Luthy, R.G. Addition of carbon sorbents to reduce PCB and PAH bioavailability in marine sediments: Physicochemical tests. Environ. Sci. Technol. 2004, 38, 5458–5464. [Google Scholar] [CrossRef] [PubMed]
  165. Uchimiya, M.; Lima, I.M.; Thomas Klasson, K.; Chang, S.; Wartelle, L.H.; Rodgers, J.E. Immobilization of heavy metal ions (Cu II, Cd II, Ni II, and Pb II) by broiler litter-derived biochars in water and soil. J. Agric. Food Chem. 2010, 58, 5538–5544. [Google Scholar] [CrossRef]
  166. Beesley, L.; Moreno-Jiménez, E.; Gomez-Eyles, J.L. Effects of biochar and greenwaste compost amendments on mobility, bioavailability and toxicity of inorganic and organic contaminants in a multi-element polluted soil. Environ. Pollut. 2010, 158, 2282–2287. [Google Scholar] [CrossRef] [PubMed]
  167. Laird, D.; Fleming, P.; Wang, B.; Horton, R.; Karlen, D. Biochar impact on nutrient leaching from a Midwestern agricultural soil. Geoderma 2010, 158, 436–442. [Google Scholar] [CrossRef]
  168. Freddo, A.; Cai, C.; Reid, B.J. Environmental contextualisation of potential toxic elements and polycyclic aromatic hydrocarbons in biochar. Environ. Pollut. 2012, 171, 18–24. [Google Scholar] [CrossRef]
  169. Mohan, D.; Pittman, C.U.; Bricka, M.; Smith, F.; Yancey, B.; Mohammad, J.; Steele, P.H.; Alexandre-Franco, M.F.; Gómez-Serrano, V.; Gong, H. Sorption of arsenic, cadmium, and lead by chars produced from fast pyrolysis of wood and bark during bio-oil production. J. Colloid Interface Sci. 2007, 310, 57–73. [Google Scholar] [CrossRef]
  170. Beesley, L.; Marmiroli, M. The immobilisation and retention of soluble arsenic, cadmium and zinc by biochar. Environ. Pollut. 2011, 159, 474–480. [Google Scholar] [CrossRef] [PubMed]
  171. Mench, M.; Bussière, S.; Boisson, J.; Castaing, E.; Vangronsveld, J.; Ruttens, A.; De Koe, T.; Bleeker, P.; Assunção, A.; Manceau, A. Progress in remediation and revegetation of the barren Jales gold mine spoil after in situ treatments. Plant Soil 2003, 249, 187–202. [Google Scholar] [CrossRef]
  172. Arán, D.; Antelo, J.; Lodeiro, P.; Macías, F.; Fiol, S. Use of waste-derived biochar to remove copper from aqueous solution in a continuous-flow system. Ind. Eng. Chem. Res. 2017, 56, 12755–12762. [Google Scholar] [CrossRef]
  173. Meier, S.; Curaqueo, G.; Khan, N.; Bolan, N.; Cea, M.; Eugenia, G.M.; Cornejo, P.; Ok, Y.S.; Borie, F. Chicken-manure-derived biochar reduced bioavailability of copper in a contaminated soil. J. Soils Sediments 2017, 17, 741–750. [Google Scholar] [CrossRef]
  174. Downie, A.; Munroe, P.; Cowie, A.L.; Van Zwieten, L.; Lau, D.M.S. Biochar as a geoengineering climate solution: Hazard identification and risk management. Crit. Rev. Environ. Sci. Technol. 2012, 42, 225–250. [Google Scholar] [CrossRef]
  175. Novak, J.M.; Lima, I.; Xing, B.; Gaskin, J.W.; Steiner, C.; Das, K.C.; Ahmedna, M.; Rehrah, D.; Watts, D.W.; Busscher, W.J.; et al. Characterization of designer biochar produced at different temperatures and their effects on a loamy sand. Ann. Environ. Sci. 2009, 3, 195–206. [Google Scholar]
  176. Yu, X.-Y.; Ying, G.-G.; Kookana, R.S. Reduced plant uptake of pesticides with biochar additions to soil. Chemosphere 2009, 76, 665–671. [Google Scholar] [CrossRef]
  177. Rogovska, N.; Laird, D.; Cruse, R.M.; Trabue, S.; Heaton, E. Germination Tests for Assessing Biochar Quality. J. Environ. Qual. 2012, 41, 1014. [Google Scholar] [CrossRef]
  178. Kuppusamy, S.; Thavamani, P.; Megharaj, M.; Venkateswarlu, K.; Naidu, R. Agronomic and remedial benefits and risks of applying biochar to soil: Current knowledge and future research directions. Environ. Int. 2016, 87, 1–12. [Google Scholar] [CrossRef]
  179. Asai, H.; Samson, B.K.; Stephan, H.M.; Songyikhangsuthor, K.; Homma, K.; Kiyono, Y.; Inoue, Y.; Shiraiwa, T.; Horie, T. Biochar amendment techniques for upland rice production in Northern Laos. 1. Soil physical properties, leaf SPAD and grain yield. Field Crop. Res. 2009, 111, 81–84. [Google Scholar] [CrossRef]
  180. Van Zwieten, L.; Kimber, S.; Morris, S.; Chan, K.Y.; Downie, A.; Rust, J.; Joseph, S.; Cowie, A. Effects of biochar from slow pyrolysis of papermill waste on agronomic performance and soil fertility. Plant Soil 2010, 327, 235–246. [Google Scholar] [CrossRef]
  181. Spokas, K.A. Review of the stability of biochar in soils: Predictability of O:C molar ratios. Carbon Manag. 2010, 1, 289–303. [Google Scholar] [CrossRef]
  182. Graber, E.R.; Tsechansky, L.; Khanukov, J.; Oka, Y. Sorption, Volatilization, and Efficacy of the Fumigant 1,3-Dichloropropene in a Biochar-Amended Soil. Soil Sci. Soc. Am. J. 2011, 75, 1365–1373. [Google Scholar] [CrossRef]
  183. Wang, J.; Pan, X.; Liu, Y.; Zhang, X.; Xiong, Z. Effects of biochar amendment in two soils on greenhouse gas emissions and crop production. Plant Soil 2012, 360, 1–12. [Google Scholar] [CrossRef]
  184. Nelson, N.O.; Agudelo, S.C.; Yuan, W.; Gan, J. Nitrogen and phosphorus availability in biochar-amended soils. Soil Sci. 2011, 176, 218–226. [Google Scholar] [CrossRef]
  185. Kammann, C.I.; Linsel, S.; Gößling, J.W.; Koyro, H.-W. Influence of biochar on drought tolerance of Chenopodium quinoa Willd and on soil–plant relations. Plant Soil 2011, 345, 195–210. [Google Scholar] [CrossRef]
  186. Glazunova, D.M.; Kuryntseva, P.A.; Selivanovskaya, S.Y.; Galitskaya, P.Y. Assessing the Potential of Using Biochar as a Soil Conditioner. IOP Conf. Ser. Earth Environ. Sci. 2018, 107, 012059. [Google Scholar] [CrossRef]
  187. Gordeev, A.; Vasserman, D.; Galitskaya, P.; Selivanovskaya, D.S. Chemical and micromorphological characteristics, biotoxicity and plant stimulating properties of the biochar obtained at different temperatures of pyrolysis. In Proceedings of the International Multidisciplinary Scientific GeoConference Surveying Geology and Mining Ecology Management, SGEM, Albena, Bulgaria, 2–8 July 2018. [Google Scholar]
  188. Lehmann, J.; Rillig, M.C.; Thies, J.; Masiello, C.A.; Hockaday, W.C.; Crowley, D. Biochar effects on soil biota–A review. Soil Biol. Biochem. 2011, 43, 1812–1836. [Google Scholar] [CrossRef]
  189. Brussaard, L.; de Ruiter, P.C.; Brown, G.G. Soil biodiversity for agricultural sustainability. Agric. Ecosyst. Environ. 2007, 121, 233–244. [Google Scholar] [CrossRef]
  190. Moore, J.C.; Berlow, E.L.; Coleman, D.C.; De Suiter, P.C.; Dong, Q.; Hastings, A.; Johnson, N.C.; McCann, K.S.; Melville, K.; Morin, P.J.; et al. Detritus, trophic dynamics and biodiversity. Ecol. Lett. 2004, 7, 584–600. [Google Scholar] [CrossRef]
  191. Solaiman, Z.M.; Blackwell, P.; Abbott, L.K.; Storer, P.P. Direct and residual effect of biochar application on mycorrhizal root colonisation, growth and nutrition of wheat. Aust. J. Soil Res. 2010, 48, 546–554. [Google Scholar] [CrossRef]
  192. Rillig, M.C.; Wagner, M.; Salem, M.; Antunes, P.M.; George, C.; Ramke, H.-G.G.; Titirici, M.-M.M.; Antonietti, M. Material derived from hydrothermal carbonization: Effects on plant growth and arbuscular mycorrhiza. Appl. Soil Ecol. 2010, 45, 238–242. [Google Scholar] [CrossRef]
  193. Killham, K.; Firestone, M.K. Salt stress control of intracellular solutes in streptomycetes indigenous to saline soils. Appl. Environ. Microbiol. 1984, 47, 301–306. [Google Scholar] [CrossRef]
  194. Covacevich, F.; Marino, M.A.; Echeverría, H.E. The phosphorus source determines the arbuscular mycorrhizal potential and the native mycorrhizal colonization of tall fescue and wheatgrass. Eur. J. Soil Biol. 2006, 42, 127–138. [Google Scholar] [CrossRef]
  195. Blackwell, P.; Krull, E.; Butler, G.; Herbert, A.; Solaiman, Z. Effect of banded biochar on dryland wheat production and fertiliser use in south-western Australia: An agronomic and economic perspective. Aust. J. Soil Res. 2010, 48, 531–545. [Google Scholar] [CrossRef]
  196. Mohamed, I.; El-Meihy, R.; Ali, M.; Chen, F.; Raleve, D. Interactive effects of biochar and micronutrients on faba bean growth, symbiotic performance, and soil properties. J. Plant Nutr. Soil Sci. 2017, 180, 729–738. [Google Scholar] [CrossRef]
  197. Rousk, J.; Brookes, P.C.; Bååth, E. Contrasting soil pH effects on fungal and bacterial growth suggest functional redundancy in carbon mineralization. Appl. Environ. Microbiol. 2009, 75, 1589–1596. [Google Scholar] [CrossRef] [PubMed]
  198. Kasozi, G.N.; Zimmerman, A.R.; Nkedi-Kizza, P.; Gao, B. Catechol and Humic Acid Sorption onto a Range of Laboratory-Produced Black Carbons (Biochars). Environ. Sci. Technol. 2010, 44, 6189–6195. [Google Scholar] [CrossRef] [PubMed]
  199. DeLuca, T.H.; MacKenzie, M.D.; Gundale, M.J.; Holben, W.E. Wildfire-Produced Charcoal Directly Influences Nitrogen Cycling in Ponderosa Pine Forests. Soil Sci. Soc. Am. J. 2006, 70, 448–453. [Google Scholar] [CrossRef]
  200. Bailey, V.L.; Fansler, S.J.; Smith, J.L.; Bolton, H. Reconciling apparent variability in effects of biochar amendment on soil enzyme activities by assay optimization. Soil Biol. Biochem. 2011, 43, 296–301. [Google Scholar] [CrossRef]
  201. Anderson, C.R.; Condron, L.M.; Clough, T.J.; Fiers, M.; Stewart, A.; Hill, R.A.; Sherlock, R.R. Biochar induced soil microbial community change: Implications for biogeochemical cycling of carbon, nitrogen and phosphorus. Pedobiologia 2011, 54, 309–320. [Google Scholar] [CrossRef]
  202. Spokas, K.A.; Baker, J.M.; Reicosky, D.C. Ethylene: Potential key for biochar amendment impacts. Plant Soil 2010, 333, 443–452. [Google Scholar] [CrossRef]
  203. Rondon, M.A.; Lehmann, J.; Ramírez, J.; Hurtado, M. Biological nitrogen fixation by common beans (Phaseolus vulgaris L.) increases with bio-char additions. Biol. Fertil. Soils 2007, 43, 699–708. [Google Scholar] [CrossRef]
  204. Kuppusamy, S.; Kumutha, K.; Santhana Krishnan, P. Influence of biochar and azospirillumapplication on the growth of maize. Madras Agric. 2011, 98, 158–164. [Google Scholar]
  205. Elmer, W.H.; Pignatello, J.J. Effect of biochar amendments on mycorrhizal associations and Fusarium crown and root rot of asparagus in replant soils. Plant Dis. 2011, 95, 960–966. [Google Scholar] [CrossRef]
  206. Grossman, J.M.; O’Neill, B.E.; Tsai, S.M.; Liang, B.; Neves, E.; Lehmann, J.; Thies, J.E. Amazonian anthrosols support similar microbial communities that differ distinctly from those extant in adjacent, unmodified soils of the same mineralogy. Microb. Ecol. 2010, 60, 192–205. [Google Scholar] [CrossRef]
  207. O’Neill, B.; Grossman, J.; Tsai, M.T.; Gomes, J.E.; Lehmann, J.; Peterson, J.; Neves, E.; Thies, J.E. Bacterial community composition in Brazilian Anthrosols and adjacent soils characterized using culturing and molecular identification. Microb. Ecol. 2009, 58, 23–35. [Google Scholar] [CrossRef]
  208. Kolb, S.E.; Fermanich, K.J.; Dornbush, M.E. Effect of charcoal quantity on microbial biomass and activity in temperate soils. Soil Sci. Soc. Am. J. 2009, 73, 1173–1181. [Google Scholar] [CrossRef]
  209. Wardle, D.A.; Nilsson, M.-C.; Zackrisson, O. Fire-derived charcoal causes loss of forest humus. Science 2008, 320, 629. [Google Scholar] [CrossRef] [PubMed]
  210. Liang, B.; Lehmann, J.; Sohi, S.P.; Thies, J.E.; O’Neill, B.; Trujillo, L.; Gaunt, J.; Solomon, D.; Grossman, J.; Neves, E.G.; et al. Black carbon affects the cycling of non-black carbon in soil. Org. Geochem. 2010, 41, 206–213. [Google Scholar] [CrossRef]
  211. Zimmermann, M.; Bird, M.I.; Wurster, C.; Saiz, G.; Goodrick, I.; Barta, J.; Capek, P.; Santruckova, H.; Smernik, R. Rapid degradation of pyrogenic carbon. Glob. Chang. Biol. 2012, 18, 3306–3316. [Google Scholar] [CrossRef]
  212. Birk, J.; Steiner, C.; Teixiera, W.; Zech, W.; Glaser, B. Microbial response to charcoal amendments and fertilization of a highly weathered tropical soil. In Amazonian Dark Earths: Wim Sombroek’s Vision; Woods, W.I., Teixeira, W.G., Lehmann, J., Steiner, C., WinklerPrins, A.M.G.A., Rebellato, L., Eds.; Springer: Dordrecht, The Netherlands, 2009; pp. 309–324. [Google Scholar]
  213. Igalavithana, A.D.; Park, J.; Ryu, C.; Lee, Y.H.; Hashimoto, Y.; Huang, L.; Kwon, E.E.; Ok, Y.S.; Lee, S.S. Slow pyrolyzed biochars from crop residues for soil metal(loid) immobilization and microbial community abundance in contaminated agricultural soils. Chemosphere 2017, 177, 157–166. [Google Scholar] [CrossRef]
  214. Ducey, T.; Novak, J.; Johnson, M. Effects of biochar blends on microbial community composition in two coastal plain soils. Agriculture 2015, 5, 1060–1075. [Google Scholar] [CrossRef]
  215. Gomez, J.D.; Denef, K.; Stewart, C.E.; Zheng, J.; Cotrufo, M.F. Biochar addition rate influences soil microbial abundance and activity in temperate soils. Eur. J. Soil Sci. 2014, 65, 28–39. [Google Scholar] [CrossRef]
  216. Makoto, K.; Tamai, Y.; Kim, Y.S.; Koike, T. Buried charcoal layer and ectomycorrhizae cooperatively promote the growth of Larix gmelinii seedlings. Plant Soil 2010, 327, 143–152. [Google Scholar] [CrossRef]
  217. Dempster, D.N.; Gleeson, D.B.; Solaiman, Z.M.; Jones, D.L.; Murphy, D.V. Decreased soil microbial biomass and nitrogen mineralisation with Eucalyptus biochar addition to a coarse textured soil. Plant Soil 2012, 354, 311–324. [Google Scholar] [CrossRef]
  218. Steiner, C.; Garcia, M.; Zech, W. Effects of charcoal as slow release nutrient carrier on N-P-K dynamics and soil microbial population: Pot experiments with ferralsol substrate. In Amazonian Dark Earths: Wim Sombroek’s Vision; Woods, W.I., Teixeira, W.G., Lehmann, J., Steiner, C., WinklerPrins, A.M.G.A., Rebelllato, L., Eds.; Springer: Dordrecht, The Netherlands, 2009; pp. 325–338. ISBN 9781402090301. [Google Scholar]
  219. Ogawa, M.; Okimori, Y. Pioneering works in biochar research, Japan. Aust. J. Soil Res. 2010, 48, 489–500. [Google Scholar] [CrossRef]
  220. Vantsis, J.T.; Bond, G. The effect of charcoal on the growth of leguminous plants in sand culture. Ann. Appl. Biol. 1950, 37, 159–168. [Google Scholar] [CrossRef]
  221. Turner, E.R. The effect of certain adsorbents on the nodulation of clover plants. Ann. Bot. 1955, 19, 149–160. [Google Scholar] [CrossRef]
  222. Pietri, J.C.A.; Brookes, P.C. Relationships between soil pH and microbial properties in a UK arable soil. Soil Biol. Biochem. 2008, 40, 1856–1861. [Google Scholar] [CrossRef]
  223. Rousk, J.; Bååth, E.; Brookes, P.C.; Lauber, C.L.; Lozupone, C.; Caporaso, J.G.; Knight, R.; Fierer, N. Soil bacterial and fungal communities across a pH gradient in an arable soil. ISME J. 2010, 4, 1340–1351. [Google Scholar] [CrossRef]
  224. Otsuka, S.; Sudiana, I.; Komori, A.; Isobe, K.; Deguchi, S.; Nishiyama, M.; Shimizu, H.; Senoo, K. Community structure of soil bacteria in a tropical rainforest several years after fire. Microbes Environ. 2008, 23, 49–56. [Google Scholar] [CrossRef]
  225. Taketani, R.G.; Tsai, S.M. The influence of different land uses on the structure of archaeal communities in Amazonian Anthrosols based on 16S rRNA and amoA genes. Microb. Ecol. 2010, 59, 734–743. [Google Scholar] [CrossRef]
  226. Khodadad, C.L.M.M.; Zimmerman, A.R.; Green, S.J.; Uthandi, S.; Foster, J.S. Taxa-specific changes in soil microbial community composition induced by pyrogenic carbon amendments. Soil Biol. Biochem. 2011, 43, 385–392. [Google Scholar] [CrossRef]
  227. Kim, J.S.; Sparovek, G.; Longo, R.M.; De Melo, W.J.; Crowley, D. Bacterial diversity of terra preta and pristine forest soil from the Western Amazon. Soil Biol. Biochem. 2007, 39, 684–690. [Google Scholar] [CrossRef]
  228. Ball, P.N.; MacKenzie, M.D.; DeLuca, T.H.; Holben, W.E. Wildfire and charcoal enhance nitrification and ammonium-oxidizing bacterial abundance in dry montane forest soils. J. Environ. Qual. 2010, 39, 1243–1253. [Google Scholar] [CrossRef]
  229. Bääth, E.; Frostegärd, A.; Pennanen, T.; Fritze, H. Microbial community structure and pH response in relation to soil organic matter quality in wood-ash fertilized, clear-cut or burned coniferous forest soils. Soil Biol. Biochem. 1995, 27, 229–240. [Google Scholar] [CrossRef]
  230. Bolster, C.H.; Abit, S.M. Biochar Pyrolyzed at Two Temperatures Affects Transport through a Sandy Soil. J. Environ. Qual. 2012, 41, 124–133. [Google Scholar] [CrossRef]
  231. Agegnehu, G.; Srivastava, A.K.; Bird, M.I. The Role of Biochar and Biochar-Compost in Improving Soil Quality and Crop Performance: A Review; Elsevier B.V.: Amsterdam, The Netherlands, 2017; Volume 119, pp. 156–170. [Google Scholar]
  232. Ogawa, M. Symbiosis of people and nature in the tropics. Farm. Japan. 1994, 28, 10–34. [Google Scholar]
  233. Shrestha, G.; Traina, S.J.; Swanston, C.W. Black Carbon’s Properties and Role in the Environment: A Comprehensive Review. Sustainability 2010, 2, 294–320. [Google Scholar] [CrossRef]
  234. Major, J.; Lehmann, J.; Rodon, M.; Goodale, C. Fate of soil-applied black carbon: Downward migration, leaching and soil respiration. Glob. Chang. Biol. 2010, 16, 1366–1379. [Google Scholar] [CrossRef]
  235. Wang, C.; Chen, D.; Shen, J.; Yuan, Q.; Fan, F.; Wei, W.; Li, Y.; Wu, J. Biochar alters soil microbial communities and potential functions 3–4 years after amendment in a double rice cropping system. Agric. Ecosyst. Environ. 2021, 311, 107291. [Google Scholar] [CrossRef]
  236. Hagemann, N.; Harter, J.; Kaldamukova, R.; Guzman-Bustamante, I.; Ruser, R.; Graeff, S.; Kappler, A.; Behrens, S. Does soil aging affect the N2O mitigation potential of biochar? A combined microcosm and field study. GCB Bioenergy 2017, 9, 953–964. [Google Scholar] [CrossRef]
  237. Blanco-Canqui, H. Does biochar improve all soil ecosystem services? GCB Bioenergy 2021, 13, 291–304. [Google Scholar] [CrossRef]
  238. Case, S.D.C.; McNamara, N.P.; Reay, D.S.; Whitaker, J. The effect of biochar addition on N2O and CO2 emissions from a sandy loam soil-The role of soil aeration. Soil Biol. Biochem. 2012, 51, 125–134. [Google Scholar] [CrossRef]
  239. Farkas, É.; Feigl, V.; Gruiz, K.; Vaszita, E.; Fekete-Kertész, I.; Tolner, M.; Kerekes, I.; Pusztai, É.; Kari, A.; Uzinger, N.; et al. Long-term effects of grain husk and paper fibre sludge biochar on acidic and calcareous sandy soils-A scale-up field experiment applying a complex monitoring toolkit. Sci. Total Environ. 2020, 731, 138988. [Google Scholar] [CrossRef]
  240. Biederman, L.A.; Harpole, W.S. Biochar and its effects on plant productivity and nutrient cycling: A meta-analysis. GCB Bioenergy 2013, 5, 202–214. [Google Scholar] [CrossRef]
  241. Qambrani, N.A.; Rahman, M.M.; Won, S.; Shim, S.; Ra, C. Biochar properties and eco-friendly applications for climate change mitigation, waste management, and wastewater treatment: A review. Renew. Sustain. Energy Rev. 2017, 79, 255–273. [Google Scholar] [CrossRef]
  242. Zhang, X.; Li, Y.; Li, H. Enhanced Bio-Immobilization of Pb Contaminated Soil by Immobilized Bacteria with Biochar as Carrier. Polish J. Environ. Stud. 2017, 26, 413–418. [Google Scholar] [CrossRef] [PubMed]
  243. Lehmann, J.; Gaunt, J.; Rondon, M. Bio-char sequestration in terrestrial ecosystems—A review. Mitig. Adapt. Strateg. Glob. Chang. 2006, 11, 403–427. [Google Scholar] [CrossRef]
  244. Roberts, K.G.; Gloy, B.A.; Joseph, S.; Scott, N.R.; Lehmann, J. Life cycle assessment of biochar systems: Estimating the energetic, economic, and climate change potential. Environ. Sci. Technol. 2010, 44, 827–833. [Google Scholar] [CrossRef]
  245. Woolf, D.; Amonette, J.E.; Street-Perrott, F.A.; Lehmann, J.; Joseph, S. Sustainable biochar to mitigate global climate change. Nat. Commun. 2010, 1, 56. [Google Scholar] [CrossRef]
  246. Kuzyakov, Y.; Bogomolova, I.; Glaser, B. Biochar stability in soil: Decomposition during eight years and transformation as assessed by compound-specific 14C analysis. Soil Biol. Biochem. 2014, 70, 229–236. [Google Scholar] [CrossRef]
  247. Kuzyakov, Y.; Subbotina, I.; Chen, H.; Bogomolova, I.; Xu, X. Black carbon decomposition and incorporation into soil microbial biomass estimated by 14C labeling. Soil Biol. Biochem. 2009, 41, 210–219. [Google Scholar] [CrossRef]
  248. Tisserant, A.; Cherubini, F. Potentials, limitations, co-benefits, and trade-offs of biochar applications to soils for climate change mitigation. Land 2019, 8, 179. [Google Scholar] [CrossRef]
  249. Hammond, J.; Shackley, S.; Sohi, S.; Brownsort, P. Prospective life cycle carbon abatement for pyrolysis biochar systems in the UK. Energy Policy 2011, 39, 2646–2655. [Google Scholar] [CrossRef]
  250. Hamedani, S.R.; Kuppens, T.; Malina, R.; Bocci, E.; Colantoni, A.; Villarini, M. Life cycle assessment and environmental valuation of biochar production: Two case studies in Belgium. Energies 2019, 12, 2166. [Google Scholar] [CrossRef]
  251. Leppäkoski, L.; Marttila, M.P.; Uusitalo, V.; Levänen, J.; Halonen, V.; Mikkilä, M.H. Assessing the carbon footprint of biochar from willow grown on marginal lands in finland. Sustainability 2021, 13, 10097. [Google Scholar] [CrossRef]
  252. Lerchenmüller, H.; Bier, H.; Hettich, L.; Gustafsson, M. European Biochar Market Report 2022/2023; European Biochar Industry: Freiburg im Breisgau, Germany, 2022. [Google Scholar]
  253. Nematian, M.; Keske, C.; Ng’ombe, J.N. A techno-economic analysis of biochar production and the bioeconomy for orchard biomass. Waste Manag. 2021, 135, 467–477. [Google Scholar] [CrossRef] [PubMed]
  254. Robb, S. Biochar making €300 to €2000 per tonne. Ir. Farmers J. 2023. Available online: https://www.farmersjournal.ie/biochar-making-300-to-2-000-per-tonne-763428 (accessed on 7 August 2023).
  255. EU Carbon Permits. Available online: https://tradingeconomics.com/commodity/carbon (accessed on 7 August 2023).
  256. Yanai, Y.; Toyota, K.; Okazaki, M. Effects of charcoal addition on N2O emissions from soil resulting from rewetting air-dried soil in short-term laboratory experiments: Original article. Soil Sci. Plant Nutr. 2007, 53, 181–188. [Google Scholar] [CrossRef]
  257. Van Zwieten, L.; Singh, B.P.; Joseph, S.; Kimber, S.; Cowie, A. Chan Biochar and emission of non-CO2 greenhouse gases from soil. In Biochar for Environmental Management: Science and Technology; Lehmann, J., Joseph, S., Eds.; Earthscan: London, UK, 2009; pp. 227–249. ISBN 978-1-84407-658-1. [Google Scholar]
  258. Spokas, K.A.; Koskinen, W.C.; Baker, J.M.; Reicosky, D.C. Impacts of woodchip biochar additions on greenhouse gas production and sorption/degradation of two herbicides in a Minnesota soil. Chemosphere 2009, 77, 574–581. [Google Scholar] [CrossRef] [PubMed]
  259. Deenik, J.L.; McClellan, T.; Uehara, G.; Antal, M.J.; Campbell, S. Charcoal volatile matter content influences plant growth and soil nitrogen transformations. Soil Sci. Soc. Am. J. 2010, 74, 1259–1270. [Google Scholar] [CrossRef]
  260. Zimmerman, A.R. Abiotic and microbial oxidation of laboratory-produced black carbon (biochar). Environ. Sci. Technol. 2010, 44, 1295–1301. [Google Scholar] [CrossRef]
  261. Hofrichter, M. Review: Lignin conversion by manganese peroxidase (MnP). Enzyme Microb. Technol. 2002, 30, 454–466. [Google Scholar] [CrossRef]
  262. Derenne, S.; Largeau, C. A review of some important families of refractory macromolecules: Composition, origin, and fate in soils and sediments. Soil Sci. 2001, 166, 833–847. [Google Scholar] [CrossRef]
  263. Das, K.C.; Garcia-Perez, M.; Bibens, B.; Melear, N. Slow pyrolysis of poultry litter and pine woody biomass: Impact of chars and bio-oils on microbial growth. J. Environ. Sci. Health-Part A Toxic/Hazardous Subst. Environ. Eng. 2008, 43, 714–724. [Google Scholar] [CrossRef]
  264. Steiner, C.; Das, K.C.; Garcia, M.; Förster, B.; Zech, W. Charcoal and smoke extract stimulate the soil microbial community in a highly weathered xanthic Ferralsol. Pedobiologia (Jena) 2008, 51, 359–366. [Google Scholar] [CrossRef]
  265. Kimetu, J.M.; Lehmann, J. Stability and stabilisation of biochar and green manure in soil with different organic carbon contents. Aust. J. Soil Res. 2010, 48, 577–585. [Google Scholar] [CrossRef]
  266. Koch, B.; Ostermann, M.; Höke, H.; Hempel, D.-C. Sand and activated carbon as biofilm carriers for microbial degradation of phenols and nitrogen-containing aromatic compounds. Water Res. 1991, 25, 1–8. [Google Scholar] [CrossRef]
  267. Quirós, M.; García, A.B.; Montes-Morán, M.A. Influence of the support surface properties on the protein loading and activity of lipase/mesoporous carbon biocatalysts. Carbon 2011, 49, 406–415. [Google Scholar] [CrossRef]
  268. Ehrhardt, H.M.; Rehm, H.J. Phenol degradation by microorganisms adsorbed on activated carbon. Appl. Microbiol. Biotechnol. 1985, 21, 32–36. [Google Scholar] [CrossRef]
  269. Kaplan, D.L.; Kaplan, A.M. Biodegradation of N-nitrosodimethylamine in aqueous and soil systems. Appl. Environ. Microbiol. 1985, 50, 1077–1086. [Google Scholar] [CrossRef]
  270. Pietikäinen, J.; Kiikkilä, O.; Fritze, H. Charcoal as a habitat for microbes and its effect on the microbial community of the underlying humus. Oikos 2000, 89, 231–242. [Google Scholar] [CrossRef]
  271. Koelmans, A.A.; Jonker, M.T.O.; Cornelissen, G.; Bucheli, T.D.; Van Noort, P.C.M.; Gustafsson, Ö. Black carbon: The reverse of its dark side. Chemosphere 2006, 63, 365–377. [Google Scholar] [CrossRef]
  272. Gomez-Eyles, J.L.; Sizmur, T.; Collins, C.D.; Hodson, M.E. Effects of biochar and the earthworm Eisenia fetida on the bioavailability of polycyclic aromatic hydrocarbons and potentially toxic elements. Environ. Pollut. 2011, 159, 616–622. [Google Scholar] [CrossRef]
  273. Noguera, D.; Rondón, M.; Laossi, K.-R.; Hoyos, V.; Lavelle, P.; de Carvalho, M.H.C.; Barot, S. Contrasted effect of biochar and earthworms on rice growth and resource allocation in different soils. Soil Biol. Biochem. 2010, 42, 1017–1027. [Google Scholar] [CrossRef]
  274. Lehmann, J.; Kuzyakov, Y.; Pan, G.; Ok, Y.S. Biochars and the plant-soil interface. Plant Soil 2015, 395, 1–5. [Google Scholar] [CrossRef]
  275. Liu, X.; Zhang, A.; Ji, C.; Joseph, S.; Bian, R.; Li, L.; Pan, G.; Paz-Ferreiro, J. Biochar’s effect on crop productivity and the dependence on experimental conditions—A meta-analysis of literature data. Plant Soil 2013, 373, 583–594. [Google Scholar] [CrossRef]
  276. Camps Arbestain, M.; Saggar, S.; Leifeld, J. Environmental benefits and risks of biochar application to soil. Agric. Ecosyst. Environ. 2014, 191, 1–4. [Google Scholar] [CrossRef]
  277. Schmalenberger, A.; Fox, A. Bacterial mobilization of nutrients from biochar amended soils. Adv. Appl. Microb. 2016, 94, 109–159. [Google Scholar] [CrossRef]
  278. Haque, A.N.A.; Uddin, M.K.; Sulaiman, M.F.; Amin, A.M.; Hossain, M.; Solaiman, Z.M.; Mosharrof, M. Rice Growth Performance, Nutrient Use Efficiency and Changes in Soil Properties Influenced by Biochar under Alternate Wetting and Drying Irrigation. Sustainability 2022, 14, 7977. [Google Scholar] [CrossRef]
  279. Mukherjee, A.; Lal, R. Biochar impacts on soil physical properties and greenhouse gas emissions. Agronomy 2013, 3, 313–339. [Google Scholar] [CrossRef]
  280. Mukherjee, A.; Lal, R.; Zimmerman, A.R. Effects of biochar and other amendments on the physical properties and greenhouse gas emissions of an artificially degraded soil. Sci. Total Environ. 2014, 487, 26–36. [Google Scholar] [CrossRef] [PubMed]
  281. Jeffery, S.; Verheijen, F.G.A.; van der Velde, M.; Bastos, A.C. A quantitative review of the effects of biochar application to soils on crop productivity using meta-analysis. Agric. Ecosyst. Environ. 2011, 144, 175–187. [Google Scholar] [CrossRef]
  282. Subedi, R.; Bertora, C.; Zavattaro, L.; Grignani, C. Crop response to soils amended with biochar: Expected benefits and unintended risks. Ital. J. Agron. 2017, 11, 161–173. [Google Scholar] [CrossRef]
  283. Vaccari, F.P.P.; Baronti, S.; Lugato, E.; Genesio, L.; Castaldi, S.; Fornasier, F.; Miglietta, F. Biochar as a strategy to sequester carbon and increase yield in durum wheat. Eur. J. Agron. 2011, 34, 231–238. [Google Scholar] [CrossRef]
  284. Baronti, S.; Alberti, G.; Vedove, G.D.; di Gennaro, F.; Fellet, G.; Genesio, L.; Miglietta, F.; Peressotti, A.; Vaccari, F.P. The biochar option to improve plant yields: First results from some field and pot experiments in Italy. Ital. J. Agron. 2010, 5, 3–11. [Google Scholar] [CrossRef]
  285. Cornelissen, G.; Martinsen, V.; Shitumbanuma, V.; Alling, V.; Breedveld, G.; Rutherford, D.; Sparrevik, M.; Hale, S.; Obia, A.; Mulder, J. Biochar effect on maize yield and soil. Characteristics in five conservation farming sites in Zambia. Agronomy 2013, 3, 256–274. [Google Scholar] [CrossRef]
  286. Yao, C.; Joseph, S.; Li, L.; Pan, G.-X.; Lin, Y.; Munroe, P.; Pace, B.; Taherymoosavi, S.; Van Zwieten, L.; Thomas, T.; et al. Developing more effective enhanced biochar fertilisers for improvement of pepper yield and quality. Pedosphere 2015, 25, 703–712. [Google Scholar] [CrossRef]
  287. Schmidt, H.; Pandit, B.; Martinsen, V.; Cornelissen, G.; Conte, P.; Kammann, C. Fourfold increase in pumpkin yield in response to low-dosage root zone application of urine-enhanced biochar to a fertile tropical soil. Agriculture 2015, 5, 723–741. [Google Scholar] [CrossRef]
  288. Nelissen, V.; Ruysschaert, G.; Müller-Stöver, D.; Bodé, S.; Cook, J.; Ronsse, F.; Shackley, S.; Boeckx, P.; Hauggaard-Nielsen, H. Short-term effect of feedstock and pyrolysis temperature on biochar characteristics, soil and crop response in temperate soils. Agronomy 2014, 4, 52–73. [Google Scholar] [CrossRef]
  289. Rogovska, N.; Laird, D.A.; Rathke, S.J.; Karlen, D.L. Biochar impact on Midwestern Mollisols and maize nutrient availability. Geoderma 2014, 230, 340–347. [Google Scholar] [CrossRef]
  290. Tammeorg, P.; Simojoki, A.; Mäkelä, P.; Stoddard, F.L.; Alakukku, L.; Helenius, J. Biochar application to a fertile sandy clay loam in boreal conditions: Effects on soil properties and yield formation of wheat, turnip rape and faba bean. Plant Soil 2014, 374, 89–107. [Google Scholar] [CrossRef]
  291. Bass, A.M.; Bird, M.I.; Kay, G.; Muirhead, B. Soil properties, greenhouse gas emissions and crop yield under compost, biochar and co-composted biochar in two tropical agronomic systems. Sci. Total. Environ. 2016, 550, 459–470. [Google Scholar] [CrossRef]
  292. Genesio, L.; Miglietta, F.; Baronti, S.; Vaccari, F.P. Biochar increases vineyard productivity without affecting grape quality: Results from a four years field experiment in Tuscany. Agric. Ecosyst. Environ. 2015, 201, 20–25. [Google Scholar] [CrossRef]
  293. Schmidt, H.-P.; Kammann, C.; Niggli, C.; Evangelou, M.W.H.; Mackie, K.A.; Abiven, S. Biochar and biochar-compost as soil amendments to a vineyard soil: Influences on plant growth, nutrient uptake, plant health and grape quality. Agric. Ecosyst. Environ. 2014, 191, 117–123. [Google Scholar] [CrossRef]
  294. Jones, D.L.; Rousk, J.; Edwards-Jones, G.; DeLuca, T.H.; Murphy, D.V. Biochar-mediated changes in soil quality and plant growth in a three year field trial. Soil Biol. Biochem. 2012, 45, 113–124. [Google Scholar] [CrossRef]
  295. Subedi, R.; Taupe, N.; Ikoyi, I.; Bertora, C.; Zavattaro, L.; Schmalenberger, A.; Leahy, J.J.; Grignani, C. Chemically and biologically-mediated fertilizing value of manure-derived biochar. Sci. Total Environ. 2016, 550, 924–933. [Google Scholar] [CrossRef] [PubMed]
  296. Marks, E.; Alcañiz, J.; Domene, X. Unintended effects of biochars on short-term plant growth in a calcareous soil. Plant Soil 2014, 385, 87–105. [Google Scholar] [CrossRef]
  297. Butnan, S.; Deenik, J.L.; Toomsan, B.; Antal, M.J.; Vityakon, P. Biochar characteristics and application rates affecting corn growth and properties of soils contrasting in texture and mineralogy. Geoderma 2015, 237–238, 105–116. [Google Scholar] [CrossRef]
  298. Singh, B.; Singh, B.P.; Cowie, A.L. Characterisation and evaluation of biochars for their application as a soil amendment. Soil Res. 2010, 48, 516. [Google Scholar] [CrossRef]
  299. Klein, B.; Bopp, M. Effect of activated charcoal in agar on the culture of lower plants. Nature 1971, 230, 474. [Google Scholar] [CrossRef]
  300. Yu, P.; Qin, K.; Niu, G.; Gu, M. Alleviate environmental concerns with biochar as a container substrate: A review. Front. Plant Sci. 2023, 14, 1–14. [Google Scholar] [CrossRef]
  301. Dumroese, R.K.; Heiskanen, J.; Englund, K.; Tervahauta, A. Pelleted biochar: Chemical and physical properties show potential use as a substrate in container nurseries. Biomass Bioenergy 2011, 35, 2018–2027. [Google Scholar] [CrossRef]
  302. Zhang, L.; Sun, X.; Tian, Y.; Gong, X. Biochar and humic acid amendments improve the quality of composted green waste as a growth medium for the ornamental plant Calathea insignis. Sci. Hortic. 2014, 176, 70–78. [Google Scholar] [CrossRef]
  303. Conversa, G.; Bonasia, A.; Lazzizera, C.; Elia, A. Influence of biochar, mycorrhizal inoculation, and fertilizer rate on growth and flowering of Pelargonium (Pelargonium zonale L.) plants. Front. Plant Sci. 2015, 6, 429. [Google Scholar] [CrossRef] [PubMed]
  304. Kim, H.S.; Kim, K.R.; Yang, J.-E.; Ok, Y.S.; Kim, W.I.; Kunhikrishnan, A.; Kim, K.-H. A melioration of horticultural growing media properties through rice hull biochar incorporation. Waste and Biomass Valorization 2017, 8, 483–492. [Google Scholar] [CrossRef]
  305. Nieto, A.; Gascó, G.; Paz-Ferreiro, J.; Fernández, J.M.; Plaza, C.; Méndez, A. The effect of pruning waste and biochar addition on brown peat based growing media properties. Sci. Hortic. 2016, 199, 142–148. [Google Scholar] [CrossRef]
  306. Steiner, C.; Harttung, T. Biochar as a growing media additive and peat substitute. Solid Earth 2014, 5, 995–999. [Google Scholar] [CrossRef]
  307. Vaughn, S.F.; Kenar, J.A.; Thompson, A.R.; Peterson, S.C. Comparison of biochars derived from wood pellets and pelletized wheat straw as replacements for peat in potting substrates. Ind. Crops Prod. 2013, 51, 437–443. [Google Scholar] [CrossRef]
  308. Frenkel, O.; Jaiswal, A.K.; Elad, Y.; Lew, B.; Kammann, C.; Graber, E.R. The effect of biochar on plant diseases: What should we learn while designing biochar substrates? J. Environ. Eng. Landsc. Manag. 2017, 25, 105–113. [Google Scholar] [CrossRef]
  309. Bonanomi, G.; Ippolito, F.; Scala, F. A “black” future for plant pathology? Biochar as a new soil amendment for controlling plant diseases. J. Plant Pathol. 2015, 97, 223–224. [Google Scholar]
  310. Graber, E.R.; Frenkel, O.; Jaiswal, A.K.; Elad, Y. How may biochar influence severity of diseases caused by soilborne pathogens? Carbon Manag. 2014, 5, 169–183. [Google Scholar] [CrossRef]
  311. Elad, Y.; David, D.R.; Harel, Y.M.; Borenshtein, M.; Kalifa, H.B.; Silber, A.; Graber, E.R. Induction of systemic resistance in plants by biochar, a soil-applied carbon sequestering agent. Phytopathology 2010, 100, 913–921. [Google Scholar] [CrossRef]
  312. Akhter, A.; Hage-Ahmed, K.; Soja, G.; Steinkellner, S. Potential of Fusarium wilt-inducing chlamydospores, in vitro behaviour in root exudates and physiology of tomato in biochar and compost amended soil. Plant Soil 2016, 406, 425–440. [Google Scholar] [CrossRef]
  313. Zwart, D.C.; Kim, S.H. Biochar amendment increases resistance to stem lesions caused by phytophthora spp. in tree seedlings. HortScience 2012, 47, 1736–1740. [Google Scholar] [CrossRef]
  314. Gravel, V.; Dorais, M.; Ménard, C. Organic potted plants amended with biochar: Its effect on growth and Pythium colonization. Can. J. Plant Sci. 2013, 93, 1217–1227. [Google Scholar] [CrossRef]
  315. Knox, O.G.G.; Oghoro, C.O.; Burnett, F.J.; Fountaine, J.M. Biochar increases soil pH, but is as ineffective as liming at controlling clubroot. J. Plant Pathol. 2015, 97, 149–152. [Google Scholar] [CrossRef]
  316. George, C.; Kohler, J.; Rillig, M.C. Biochars reduce infection rates of the root-lesion nematode Pratylenchus penetrans and associated biomass loss in carrot. Soil Biol. Biochem. 2016, 95, 11–18. [Google Scholar] [CrossRef]
  317. Jie, L.; Ding, Y.; Ma, L.; Gao, G.; Wang, Y. Combination of biochar and immobilized bacteria in cypermethrin-contaminated soil remediation. Int. Biodeterior. Biodegrad. 2017, 120, 15–20. [Google Scholar]
  318. Chen, B.; Yuan, M.; Qian, L. Enhanced bioremediation of PAH-contaminated soil by immobilized bacteria with plant residue and biochar as carriers. J. Soils Sediments 2012, 12, 1350–1359. [Google Scholar] [CrossRef]
  319. Zhang, H.; Tang, J.; Wang, L.; Liu, J.; Gurav, R.G.; Sun, K. A novel bioremediation strategy for petroleum hydrocarbon pollutants using salt tolerant Corynebacterium variabile HRJ4 and biochar. J. Environ. Sci. (China) 2016, 47, 7–13. [Google Scholar] [CrossRef]
  320. Wang, T.; Sun, H.; Ren, X.; Li, B.; Mao, H. Evaluation of biochars from different stock materials as carriers of bacterial strain for remediation of heavy metal-contaminated soil. Sci. Rep. 2017, 7, 12114. [Google Scholar] [CrossRef]
Figure 1. Carbon turnover in the production and application of biochar to soils.
Figure 1. Carbon turnover in the production and application of biochar to soils.
Agriculture 13 02003 g001
Figure 2. Classification of biochar modification approaches.
Figure 2. Classification of biochar modification approaches.
Agriculture 13 02003 g002
Figure 3. SEM images from biochar prepared (a)—from plant residues, (b)—from chicken manure, and (c)—from sewage sludge.
Figure 3. SEM images from biochar prepared (a)—from plant residues, (b)—from chicken manure, and (c)—from sewage sludge.
Agriculture 13 02003 g003
Figure 5. Van Krevelen diagram for the natural carbonization process [135].
Figure 5. Van Krevelen diagram for the natural carbonization process [135].
Agriculture 13 02003 g005
Table 1. Classification of sources for biochar production.
Table 1. Classification of sources for biochar production.
Biomass GroupBiomass Sub-Groups
Plant biomass
  • Wood biomass (sawdust, lumps, stems, branches, foliage)
  • Grass biomass (plant planting for biomass production: alfalfa, arundo, bamboo, banana, brassica, cane, cynara, miscanthus, switchgrass, timothy, etc.)
  • Aquatic biomass (marine or freshwater algae; macroalgae)
Organic waste
  • Animal manure (chicken manure, cow manure, horse manure, pig manure, turkey poultry manure, etc.)
  • Crop and food processing waste (straws, spoiled fruits, shells, husks, hulls, pits, pips, grains, spoiled seeds, coir, stalks, cobs, kernels, bagasse, spoiled food, fodder, pulps, cakes, etc.)
  • Municipal wastes (Sewage sludge, organic fraction of municipal solid waste, paper-pulp sludge, waste papers, chip-board, fiberboard, plywood, wood pallets and boxes)
Table 2. Composition of lignocellulosic biomass of various origin.
Table 2. Composition of lignocellulosic biomass of various origin.
Biomass SourceComposition, %References
CelluloseHemicelluloseLigninExtractivesAsh
Soft wood41242820.4[29]
45–5025–3525–35--[30]
Hard wood39352030.3[29]
40–5524–4018–25--[30]
Pine bark341634142[29]
Wheat straw402817117[31,32]
Rice husk3025121816[33]
Hazelnut412726-3[34]
Orange peels1462-1.5[35]
Sugarcane bagasse353325-4[36]
Rice straw322418-1.2[37]
Banana60–656–85–10-1.2[38]
Newspaper40–5525–3018–25--[39]
Corn cobs453515--[40]
Sponge gaurs fibers671715--[38]
Table 3. Chemical composition of algae, sludge and manure biochar feedstocks.
Table 3. Chemical composition of algae, sludge and manure biochar feedstocks.
Biomass SourceChemical Content (wt.%)Proximate Analysis, wt.%Reference
ProteinsLipidsCarbohydratesMoistureVolatile MatterFixed CarbonAsh
Chlorella53.81.037.1076.414.59.1[47]
Cladophora glomerata27.85.332.44.546.314.734.5[48]
Sewage sludge26.327.322.93.653.96.236.3[49]
Sewage sludge---16.864.410.824.8[50]
Chicken manure31.62.134.5---15.9[51]
Swine manure229.139.1---15.1
Cow manure18.18.752.6---12.0
Cattle manure----64.620.714.7[22]
Chicken manure----64.814.920.2
Table 4. Biochar yield under various production conditions and feedstock used.
Table 4. Biochar yield under various production conditions and feedstock used.
Pyrolysis TypeFeedstockPyrolysis Temperature, °CBiochar Yield, %Ref.
SlowRice straw30050[52]
50039
70036.5
SlowSwitchgrass40048[52]
60025
SlowCanola straw35024[53]
Rice straw35033
Soybean straw35032.5
Pea straw35032
SlowPoultry litter30060[54]
40052
50048
60046
SlowMagnolia leaves30062[55]
Apple wood60025
Spotted gum wood40051
SlowBlack locust wood30042[56]
50024
SlowPine chips55030[57]
SlowSawdust55028[58]
FastPitch pine wood chips30061[59]
FastPine sawdust40055[60]
80018
TorrefactionCoffee ground30081[53]
Microalgae residues 56
Table 5. Chemical properties of biochar produced from different feedstocks by pyrolysis (Nguyen et al., 2010) [137].
Table 5. Chemical properties of biochar produced from different feedstocks by pyrolysis (Nguyen et al., 2010) [137].
FeedstockOak Wood Corn StoverPoultry Litter
Pyrolysis temperature350600350600350600
pH4.806.389.399.429.6510.33
CEC, mmol/kg29476419252112159
C, %758860712924
C/N45548951661525
P, mg/kg12291890211421,25623,596
Fixed C387140601.60.1
H/C0.550.330.750.390.570.18
O/C0.200.070.290.100.410.62
Table 6. Agrochemical and Ecotoxicological Consequences of Biochar Application to Soils.
Table 6. Agrochemical and Ecotoxicological Consequences of Biochar Application to Soils.
BiocharApplicationEffectRef.
PositiveNegative
Wheat straw0–40 t/ha + 300 kg N/haIncrease in pH, C, N, decrease in bulk density, decrease in N2O emissionIncrease in methane emission (34–41%) when applying 40 t/ha[179]
Eucalyptus chips0–90 g/kg of soil + rhizobium cultureIncrease in biological nitrogen fixation, pH, bioavailability of soil B, Mo, R, Ca, P at 60 g/kgDecrease in available nitrogen and yield when applying 90 g/kg of biochar[138]
Paper production waste10 t/ha + Nutricote fertilizer 12.5 g/250 g of soil (15% N, 4.7% P, 8.9% K)Increase in organic carbon, pH, exchangeable Ca, and in total carbonDecrease in the yields of wheat and radishes[180]
Sawdust20–60% of biochrIncreased sorption of atrazine and acechlor by 5% biocharIncrease in the yield of N2O and CH4 when applying 20% of biochar[181]
Wood13–52 t/haDecrease in the efficiency of herbicides when the dose of biochar increases[182]
Eucalyptus wood1% of coalDecrease in bioavailability of insecticides, and in the accumulation of herbicides by plants by 10–25%Increase in the resistance of herbicide residues in soil fertilized with biochar[176]
Eucalyptus wood0.5% of coalDecreased bioavailability of chlorantraniliproleBiochar influence on pesticide bioavailability depended on soil type[183]
Corn cobsFertilizer from 2–20 g/kg of coal + N (0–100 mg/kg) + P (0–20 mg/kg), biochar 100 and 200 t/haAdditional nitrogen application is required to increase the crop yield[184]
Peanut husk100 and 200 t/haIncreased yield and drought tolerance Chenopodium quinoaA positive effect is achieved only at the maximum application rate (200 t/ha)[185]
Table 7. Possible mechanism of biochar’s influence on microbial abundance in soil [188].
Table 7. Possible mechanism of biochar’s influence on microbial abundance in soil [188].
Mechanism of InfluenceRhizobiaOther BacteriaMycorrhizal FungiOther Fungi
Protection from grazers0(+)(+)(+)
Improved hydration++?? or ±
Greater P, N Ca, Mg, K availability++--
Greater micronutrient availability++-?
Higher pH++ncnc
Lower pH--ncnc or -
Sorption of signaling compounds? or -???
Greater N availability (also through sorption of phenolics and increased nitrification-+ or -ncnc
Sorption of microorganismsnc?ncnc
Biofilm formation++??
Sorption of inhibitory compounds?+??
Sorption of dissolved organic matter??nc?
‘+’—relative abundance increase, ‘-‘—abundance decrease ‘0′—no influence observed, ‘?’—influence not considered; ‘±’—multidirectional changes; ‘nc’—influence not calculated; parentheses—weak circumstantial evidence.
Table 8. Average yield changes as a percentage of control for various biochar applications.
Table 8. Average yield changes as a percentage of control for various biochar applications.
Dose of Biochar, t/haThe Mean Values and the Range Corresponding to the 95% Confidence Interval
135−26–67
10015–65
67.5−16–45
65−10–50
5016–35
40−8–38
2510–38
22−13–18
20−15–40
16−28–22
14−14–28
11−11–17
105–17
8−5–17
6−3–16
5.5−40–50
5−56–7
4−32–22
3−5–15
1.5−7–12
Table 9. Effect of biochar application on the productivity of various crops (according to Subedi et al., 2017) [282].
Table 9. Effect of biochar application on the productivity of various crops (according to Subedi et al., 2017) [282].
Biochar SourcePyrolysis ConditionsApplication DoseTermCropEffectRef.
Mixed wood500 °C30 and 60 t/ha2 seasonsHard wheat+30% (grain)[283]
Mixed wood500 °C10 t/ha1 seasonHard wheat, corn+10% (wheat), +(6–24)% corn[284]
30 and 60 t/ha8.5 weeksL. perenne+(20–29)% (biomass)
100 and 120 t/ha8.5 weeksL. perenne−(10–20)% (biomass)
Corn heart, soft wood400 °C4 t/ha2 seasonsCorn+(233–322)%[285]
Wheat straw450 °C, enriched with minerals670 kg/ha15 weeksGreen pepper+(16–16)% (fruits)[286]
Eupatorium adenophorum680–700 °C750 kg/ha17 weeksPumpkin +(85–300)%[287]
Eucalyptus600 °C, activated1.1 and 5.44 t/ha1 harvest cycleSweet corn-[288]
Hard wood500–575 °C (hydrothermal)0–96 t/ha1 harvest cycleCorn+(11–55)%[289]
Spruce and pine chips550–600 °C5 and 10 t/ha3 yearsHorse beans, turnips and wheat-[290]
Willow 600 °C, decomposted10 t/ha (biochar), 25 t/ha (biochar and compost)1 harvest cycle for each cropBananas, papaya−(18–24)% bananas,
-papaya
[291]
Fruit tree branches500 °C22 and 44 t/ha4 yearsVine +(16–66)%[292]
Hard wood500 °C8 t/ha3 yearsVine-[293]
Mixed wood chips450 °C25 and 50 t/ha3 yearsCorn (year 1) and hay grass (years 2 and 3)9 (corn), +(13–32)% (hay)[294]
Table 10. Study of biochar as a substitute for peat in a nutrient medium.
Table 10. Study of biochar as a substitute for peat in a nutrient medium.
CropDose of Positive Effect (% of Biochar)Dose of Negative Effect (%of Biochar)Ref.
Gaillardia (Gaillardia spp.)2550[301]
Calathea (Calathea insignis)20–35 [302]
Pelargonium zonale (Pellargonium zonale)3070[303]
Kale (Brassica oleracea L. var. acephala)1–5 [304]
Lettuce (L. saliva)50–75 [305]
Sunflower (Helianthus annuus)25–75100[306]
Tomatoes (Solanium lycopesicum)5 [307]
Table 11. Stimulation of plant diseases by biochar introduced into the nutrient medium.
Table 11. Stimulation of plant diseases by biochar introduced into the nutrient medium.
Pathogen PlantBiochar Raw MaterialBiochar DosisRef.
Botrytis cinereaC. annumCitrus wood3%[311]
Fusarium oxysporum
f.sp lycopersici
L. esculatumWastes of processing wood and green parts of trees3%[312]
Phytophthora
cinnanomi
Quercus rubraWood>5%[313]
Pythium ultimumC. annum,
Ocimum basilicum
Spruce bark50% (oб.)[314]
Plasmodiophora
brassica
Brassica rapaMiscanthus0.5%[315]
Pratylenchus penetransDaucus carotaPine (wood and bark), wheat husk0.8%[316]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Kuryntseva, P.; Karamova, K.; Galitskaya, P.; Selivanovskaya, S.; Evtugyn, G. Biochar Functions in Soil Depending on Feedstock and Pyrolyzation Properties with Particular Emphasis on Biological Properties. Agriculture 2023, 13, 2003. https://doi.org/10.3390/agriculture13102003

AMA Style

Kuryntseva P, Karamova K, Galitskaya P, Selivanovskaya S, Evtugyn G. Biochar Functions in Soil Depending on Feedstock and Pyrolyzation Properties with Particular Emphasis on Biological Properties. Agriculture. 2023; 13(10):2003. https://doi.org/10.3390/agriculture13102003

Chicago/Turabian Style

Kuryntseva, Polina, Kamalya Karamova, Polina Galitskaya, Svetlana Selivanovskaya, and Gennady Evtugyn. 2023. "Biochar Functions in Soil Depending on Feedstock and Pyrolyzation Properties with Particular Emphasis on Biological Properties" Agriculture 13, no. 10: 2003. https://doi.org/10.3390/agriculture13102003

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop