Next Article in Journal
Does Participation in the “Grain for Green Program” Change the Status of Rural Men and Women? An Empirical Study of Northeast China
Next Article in Special Issue
Insights into the Roles of Surface Functional Groups and Micropores in the Sorption of Ofloxacin on Banana Pseudo-Stem Biochars
Previous Article in Journal
Understanding Factors That Influence Pest Risk in Olive Production
Previous Article in Special Issue
Enzyme Activity and Dissolved Organic Carbon Content in Soils Amended with Different Types of Biochar and Exogenous Organic Matter
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Effect of Different Biochar Characteristics on Soil Nitrogen Transformation Processes: A Review

by
Donna Davys
1,*,
Francis Rayns
1,
Susanne Charlesworth
1 and
Robert Lillywhite
2
1
Centre for Agroecology, Water and Resilience, Coventry University, Coventry CV8 3LG, UK
2
Warwick Crop Centre, School of Life Science, The University of Warwick, Coventry CV35 9EF, UK
*
Author to whom correspondence should be addressed.
Sustainability 2023, 15(23), 16446; https://doi.org/10.3390/su152316446
Submission received: 17 October 2023 / Revised: 13 November 2023 / Accepted: 20 November 2023 / Published: 30 November 2023
(This article belongs to the Special Issue Sustainable Development and Application of Biochar)

Abstract

:
For the last 30 years, interest has focused on biochar and its potential to store carbon in soil to mitigate climate change whilst improving soil properties for increased crop production and, therefore, could play a critical role in both agricultural sustainability and broader environmental aims. Biochar, a carbonaceous product, is formed from organic feedstock pyrolysised in the absence of air and, therefore, is a potential means of recycling organic waste. However, different feedstock and pyrolysis conditions result in a biochar with a range of altered characteristics. These characteristics influence nitrogen transformation processes in soil and result in the metabolism of different substrates and the formation of different products, which have different effects on agricultural yield. This paper reviews how the production of biochar, from varying feedstock and pyrolysis conditions, results in different biochar characteristics that influence each stage of the nitrogen cycle, namely processes involved in fixation, assimilation, mineralisation and denitrification. The nitrogen cycle is briefly outlined, providing a structure for the following discussion on influential biochar characteristics including carbon composition (whether recalcitrant or rapidly metabolisable), mineral composition, surface area, porosity, cation exchange capacity, inhibitory substances and pH and so on. Hence, after the addition of biochar to soil, microbial biomass and diversity, soil porosity, bulk density, water-holding capacity, cation exchange capacity, pH and other parameters change, but that change is subject to the type and amount of biochar. Hence, products from soil-based nitrogen transformation processes, which may be beneficial for plant growth, are highly dependent on biochar characteristics. The paper concludes with a diagrammatic summation of the influence of biochar on each phase of the nitrogen cycle, which, it is hoped, will serve as a reference for both students and biochar practitioners.

Graphical Abstract

1. Introduction

Every year, 2 billion tonnes of solid waste are generated globally [1], 30 billion tonnes of carbon dioxide (CO2) [2] are emitted and 12 million hectares of agricultural land are lost to soil degradation, leading to a potential loss of 20 million tonnes of grain per annum [3]. This, as well as an ever-growing food and energy requirement as tastes shift towards a more technologically driven lifestyle, means that further stress is placed on Earth’s already depleted and increasingly polluted resources [4,5]. Hence, there is now a focus on technologies that can reuse waste, reduce emissions, mitigate climate change, generate energy and support food production. One such technology is pyrolysis and the production of biochar. Pyrolysis generally refers to a reaction in an inert environment [6]. In this case, it is the thermal degradation of organic material (wood, waste or energy crops) in the absence of oxygen [7]. As well as the production of energy, the co-product of pyrolysis is a carbonaceous product referred to as biochar.
Since 1993, interest in biochar has focused on whether it can be used to store carbon in soil whilst improving soil properties for improved crop production [8,9]. Carbon storage in soils is seen as a means to mitigate greenhouse gas (GHG) emissions through the sequestration of CO2. Woolf et al. [10] analysed what level of carbon sequestration, through the production of dedicated biomass for biochar, might be practical to mitigate climate change, without compromising food security, habitat or soil conservation. They estimated that a maximum of 1.8 Pg CO2-C equivalent (CO2-Ce) per year (or 12% of current anthropogenic emissions) might be achievable, although estimates range from 0.7 to 1.8 Gt CO2 eq y−1 [11].
Biochar improves the fertility of some soils through the direct provision of nutrients such as nitrogen, phosphorus and potassium. Table 1 provides a selected range of biochar nutrient levels. Biochar can also have a liming effect, influencing pH such that nutrients become more available [12,13]. Most biochars have a high pH (when compared to soil), which results from their ash content and base cations, but also due to intrinsic alkaline organic functional groups [14]. However, these values are governed by both feedstock and pyrolysis conditions. For instance, one study found that the pH of herbaceous biochars was two units higher (9.4) than woody biochars (7.4) due to higher concentrations of ash [15]. pH was also found to be higher in biochars from leguminous feedstock (9.02 to 10.35) than in non-leguminous feedstock (8.00 to 9.24) [16]. Here, carbonates and organic anions of carboxyl and phenolic groups were the main alkalis but, again, this varied with feedstock. The influence of biochar on soil pH is key because it has been shown to improve gross mineralisation rates, immobilisation rates and heterotrophic nitrification and have an overall positive effect on soil nitrogen retention [17].
Hence, these differences in biochar qualities result in different effects on crop yield. For instance, corncob biochar on infertile acidic soil reduced maize yield, which may have resulted from the biochar’s high volatile matter content and bioavailable carbon. This labile carbon fuelled an increase in micro-organism growth, drawing nutrients from the soil and resulting in nitrogen immobilisation [18]. The availability of nitrogen for crop growth is a key agronomic parameter. Nitrogen is an essential plant nutrient and a key element in the development of organic molecules including amino acids, amino sugars, proteins, deoxynucleic acid (DNA), chlorophyll, etc. Nitrogen is very reactive and therefore has a very complex cycle through the soil system, sometimes transforming quickly between inorganic forms as gasses such as NH3, N2, N2O, and NO, in ionic form in soil, NH4+, NO2 and NO3, and organic forms. Soils vary considerably in their total nitrogen concentrations from around 0.06% to 0.30%, and more than 90% of this is in organic form [19].
Biochar has been reported to reduce nitrogen leaching losses [20]. However, these ameliorative effects can vary depending on the carbon content of the amended soil, the biochar’s carbon content, soil textural differences, soil water-holding capacity and the soil micro-organism community amongst other parameters [21,22,23,24,25]. Indeed, in their global meta-analysis of the effect of biochar on yield in temperate and tropical systems, Jeffery et al. [26] found that, on average, there was no beneficial effect on crop yields in temperate latitudes, but there was an average 25% increase in yield in the tropics. This was because in low-pH, low-fertility soils, the liming and fertilisation effect of biochar was enough to increase yield [26].
However, biochar’s influence on the native micro-organism community can exacerbate nitrogen losses from soil systems, depleting nitrogen from soil but also increasing pollutant release. For example, whilst several studies have reported that the incorporation of biochar into different soils reduced N2O emissions [27,28,29,30,31], other studies have reported that biochar incorporation results in an increase in N2O emissions [32,33]. However, the magnitude of change in N2O emission because of biochar incorporation is dependent upon experimental conditions, biochar type, application rate, soil properties, and chemical forms of added fertilizer [30]. Nevertheless, as N2O has a global warming potential of 298 times that of CO2 [34], research on techniques to minimise emissions is obviously critical.
It would appear then that biochar could prove to be a sustainable amendment for agricultural use, capable of both GHG emission reduction and soil enhancement. However, as noted above by Woolf et al. [10], it is critical that its feedstock comes from sustainable sources and does not compromise a nation’s ability to grow its own food and its food security by growing feedstock on productive land [11]. In addition, this review considers how the application of biochar to soil may influence nitrogen transformation in terms of when and where soil nitrogen is partitioned. This is a key sustainability issue because it either exacerbates or mitigates the effect and level of nitrogenous inputs and outputs to and from the soil system acting as either pollutants (for instance, in the case of GHG emission) or sources of agronomically beneficial nitrogen.

2. Phases of the Nitrogen Cycle

Here, we briefly outline the nitrogen cycle and components upon which biochar may act. This forms the basis for the following summary of recent and older investigations on the influence of different biochar characteristics on nitrogen transformations in soil. Both agronomically beneficial and environmentally damaging issues are discussed. Finally, the findings are restructured so that all influential parameters at each stage of the nitrogen cycle, here defined as fixation (of atmospheric N2), assimilation, mineralisation and denitrification, are elucidated. We hope this review will provide a reference for students, researchers and soil remediation practitioners on the influence of biochar on soil nitrogen transformation processes.
The soil system is open to the atmosphere and nitrogen cycles between them. The first input of nitrogen from the atmosphere is either through lightening, deposition or biological nitrogen fixation (BNF) (Figure 1). This is an important phase of the nitrogen cycle, capturing N2 and transforming it to NH4+. This process fixes a considerable level of nitrogen year on year. For instance, prior to human intervention, BNF is estimated to have provided 58 Tg N yr−1 worldwide, and it still plays a critical role in maintaining soil fertility [35]. It uses the soil’s natural resources (native micro-organisms) and is, therefore, a free and potentially ecologically sound means of crop fertilisation.
BNF is carried out by several groups of prokaryotes, including Azotobacter, Azospirillum, Rhizobium and Bradyrhizobium [36]. These are either free living in the soil or form symbiotic relationships with plants (around or on the root surface) and both utilise enzymatic processes to catalyse the conversion of N2 into NH3. This conversion to ammonia (or ammonification) is then followed by a process of nitrification to nitrites and nitrates. This process is undertaken by both heterotrophic and autotrophic micro-organisms with the bacterium, such as Nitrosomonas, converting NH3 into NO2, and Nitrobacter converting NO2 to NO3 [36]. It is in these forms that plants and micro-organisms can assimilate NH3 and NO3 into their structures and the nitrogen then exists in organic form. Upon their death, or through excreta, these forms are then mineralised back again into inorganic form.
They are mineralised by micro-organisms that derive energy from the oxidation of organic nitrogen to NH4+. This is then available to be assimilated and incorporated into amino acids or used for other metabolic purposes. If micro-organisms produce NH4+ in excess of their own requirements, the surplus is excreted into, in this case, the soil, and is available for assimilation by plants, or as the substrate for nitrification [37].
Finally, denitrification releases the N2 originally acquired from the atmosphere by nitrogen-fixing bacteria. Again, it is a process facilitated by micro-organisms where NO3 is reduced through a series of nitrogen oxide products of decreasing oxidation states until molecular nitrogen (N2) is all that remains. The process is of particular interest to climate change researchers as some of the intermediate products, nitric oxide and N2O, are GHGs [38].

3. The Influence of Key Biochar Parameters on the Nitrogen Cycle

Biochar influences the composition, biomass and enzyme activity of micro-organism communities, including those with nitrogen-transforming capabilities. The mechanisms include the provision of recalcitrant and metabolisable carbon, mineral ions, its sorption phenomena, change in pH, as well as physical properties such as pore structure and surface area [7,15,39,40,41]. The change in these processes results in the production of more or less substrates and products in the nitrogen cycle.

3.1. The Influence of Recalcitrant and Metabolisable Carbon

Critical to biochar’s ability to store carbon in soil are its aromatic rings, which are highly stable. Aromatic rings include six carbon structures such as benzene, which have strong chemical bonds very resistant to attack by micro-organisms [42]. It is this resistance that confers recalcitrance to biochar stored in soil and, hence, its role as a climate change mitigation tool. However, biochar contains other carbon compounds, such as aliphatic carbon chains. These structures have weaker chemical bonds easily broken by micro-organism enzymes and are thus used as an energy source for respiration and therefore result in the ultimate release of CO2. Hence, although biochar has a carbon content that can range from 50 to 95%, some of this is metabolically available and will therefore influence micro-organism growth in the short term [43,44]. The amount of metabolically available carbon depends on the biochar feedstock and pyrolysis condition. For instance, where the production of biochar involves a shorter duration in the reactor, there may be an incomplete conversion of the feedstock biomass to more recalcitrant forms, leaving a greater fraction of unconverted cellulosic and hemicellulosic forms [45]. Any carbohydrates that remain after the pyrolysis process are rapidly metabolized by soil micro-organisms and support the growth of heterotrophic populations creating changes in the relative abundance of different families [39,46,47]. However, the consequence of this growth and change in abundance of different micro-organism populations varies with the phase of the nitrogen cycle. For instance, biochar addition creates a change in abundance of Bradyrhizobiaceae and Hyphomicrobiaceae families with an increase of up to 11% relative abundance compared to control [39]. Yet, not all metabolically available carbon positively influences BNF rates. For instance, Güereña et al. [48] found that biochar with a reduced volatile matter component considerably increased micro-organism responses compared with unaltered biochar and, therefore, concluded that some volatile matter components were toxic to Rhizobium.
In general, the effect of biochar on soil micro-organism mass (and therefore soil respiration) is not well understood [49]. For example, some studies have found a 30% degradation of carbon in the first 30 years and very little thereafter (over a 100-year timeframe) [50], whereas others estimated a loss of 47% C within a 50-year timeframe [51]. Confounding factors include whether the response is dictated by the metabolisable fraction of the biochar or whether the native micro-organisms have the necessary enzymes to mineralise that available carbon [41,46].
There are potentially negative agronomic consequences resulting from the provision of metabolisable carbon. Although biochar may improve nutrient assimilation by micro-organisms, there may be a consequent decrease in assimilation by crop plants. For instance, biochars with a high aliphatic concentration are more easily degraded by soil organisms, causing an increase in their growth rates and consequent short-term assimilation and immobilisation of nitrogen, impeding supply to crop plants and reducing yield [18,52].
Conversely, biochar also stimulates the growth of micro-organisms able to mineralise recalcitrant nitrogen found in soil organic matter [53]. For instance, one N15 isotope tracing study using maize biochar found that nitrogen mineralised to NH4+ came from soil organic matter, which was otherwise generally resistant to micro-organism attack [28]. They hypothesised that this was due mainly to improved metabolisable carbon availability supplied by biochar pyrolysised at a low temperature. The subsequent elevated growth in micro-organism population could not be met with available nitrogen sources in the soil resulting in the mineralisation of recalcitrant soil organic matter [28,39,54]. Yet other studies have found no effect on gross mineralisation [55,56]. It is likely that these divergent effects are dependent on two main attributes. Firstly, the C:N ratio of the added biochar, and secondly, the nature of the native soil micro-organism community and its genetic capability to mineralise recalcitrant nitrogen. Contrasting results can be obtained if soils have different levels of nitrogen availability because soil organisms will only mineralise recalcitrant nitrogen if readily available nitrogen is low. On the other hand, in nitrogen-rich biochars, such as those from manure and maize, the excess nitrogen allows mineralisation rates to increase [28,57].
A key phase within nitrogen transformations is denitrification, because, if incomplete, it can lead to the emission of N2O. In their work with wood and poultry manure biochars, Singh et al. [22] found that, generally, biochar-amended soils produced significantly lower N2O emissions than their respective controls but its efficacy in this regard was soil-dependent. They subjected two contrasting soils with different biochar treatment rates to wet–dry cycles. By the third cycle, they found that all biochar treatments consistently decreased N2O emissions, cumulatively by 14 to 73% from the Alfisol and by 23 to 52% from the Vertisol, relative to their controls. However, some biochars produce higher N2O emissions [22,30,58]. For example, a poultry manure biochar produced at 400 °C had the highest metabolisable nitrogen content. It appears that high metabolisable carbon and nitrogen combine to promote the activity of soil denitrifiers and therefore high N2O emission [22]. Hence, where biochar provides a high metabolisable carbon content, either with or in the presence of a high nitrogen content, this can result in increased N2O emission. In this study, it was also found that N2O and CO2 emission were initially positively related, and this was consistent with other studies (for example, [59]), but that this relationship disassociates due to a rapid increase in micro-organism growth supported by the provision of metabolisable carbon after two or three days. Furthermore, the addition of biochar with metabolically available carbon leads to temporary anaerobic conditions and the enhancement of N2O-reductase resulting in a decrease in N2O release (and an increase in N2) [22,59,60]. This suggests that the initial peak of N2O release is short-lived [61]. In their review of the potential for biochar as an abatement technology for GHG, [62] also found contradictory results with much depending on biochar type.
Biochar influences the cycles of carbon already in the soil, most notably organic matter and humus. Organic matter plays a vital role in the functioning of a healthy soil. Derived from the biomass of both plants and animals (for example, roots and manure), organic matter has a high CEC, helps retain water in the soil and provides soil structure. Organic matter decays over time, releasing its nutrients. When the level of decay slows significantly or stops, and the organic matter becomes stable, it is referred to as humus. Humic acids comprise a large group of chemicals, which perform a vital role in soil health contributing to soil moisture and nutrient retention, as well as the bulk density of soil [63]. There is now some evidence to suggest that biochar serves to protect humic acids from decomposition [64]. It is thought that this may be due to a physical ‘shielding’ by biochar of humic acids as micro-organisms cannot access the smallest biochar pores [41].
In some studies, biochar promoted the degradation of organic matter, for instance, with newly incorporated plant residues. Using 14C-labeled maize residue, Awad [65] reported that biochar in both sandy and sandy loam soils stimulated soil micro-organisms, causing a significant increase in the presence of extracellular enzyme activities and consequently faster plant residue decomposition. The decomposition of plant residue was more pronounced in sandy soil, where it accounted for 23% of 14C input, whereas in sandy loam soil, increased plant residue decomposition did not exceed 14% compared to untreated soils [65]. This may have been due to differences in the nature and quantity of pre-existing soil organic matter in both soils.

3.2. The Influence of Mineral Ions, Provision, Availability and Uptake

Biochar offers the opportunity to improve soil fertility through the direct provision of mineral ions. The quantity and type of minerals depend on pyrolysis conditions and feedstock. Hardwood biochars tend to have lower mineral contents (5–10%), whereas chicken litter waste can have an ash content of up to 64% [66]. The presence of mineral ions can have a significant effect on fixation rates. For instance, Rondon et al. [31] found that biochar significantly enhanced biological fixation by Phaseolus vulgaris and accounted for the increase in fixation rate through the increased availability of molybdenum and boron provided by biochar. They noted that the improvement in BNF, as well as the increase in biomass, of Phaseolus vulgaris (5–39%) was above those normally provided by recommended commercial fertiliser applications. Similarly, in Glycine max L., the enhancement of nodulation and BNF response to carbonized organic materials was due to an increase in available sulphur [67].
Even low-nutrient biochar has the potential to elevate nutrient availability. In their greenhouse experiment with seven different biochars, [48] stripped the biochars of mineral or volatile matter, or both, and left some untreated. The untreated biochar soil treatment planted with Phaseolus vulgaris resulted in an increase of 2126% in nitrogen fixation over the control average (as well as a 262% increase in shoot biomass, a 164% increase in root biomass and a 3575% increase in nodule biomass). The stripped biochar revealed that simple mineral nutrients provided by the biochar were only slightly responsible for these increases. For instance, although the amount of nitrogen fixed was significantly correlated with plant phosphorus uptake, it was not correlated with biochar phosphorus addition but rather improved phosphorus nutrition resulting from 360% greater mycorrhizal colonization with biochar additions.
Mycorrhizal root colonisation and hyphal responses to different biochars can vary substantially [68]. This may be due to different responses to a range of volatile matter compounds from contrasting biochars with mycorrhizal response being governed by carbohydrate availability or due to biochar providing physical protection from fungal grazers or the facilitation of root and hyphal exploration, facilitating improved access to nutrients for crop plants [48,68,69].
The provision of nutrients by biochar influences several different assimilation mechanisms. For instance, where biochar provides an increased cation concentration, plant water uptake increases due to the net increase in accumulated osmotically active ions such as potassium. This is key because it improves drought tolerance [70]. In addition, biochar appears to increase tap-root growth (and potentially fine root mass), which would increase water uptake from biochar pores [71]. Evidence for this improved water status with biochar has been demonstrated through lower proline (an amino acid associated with cell osmotic adjustment in leaves) concentrations and higher osmotic values in leaves, which may reflect an increased tolerance to drought conditions [72]. In addition, biochar may not only provide more nutrients but may reduce the leaching of nutrients from the soil, thereby maintaining availability [73]. One further mechanism is a biochar-mediated reduction in transpiration. For instance, one study found that biochar-amended plants produced larger leaves, and the plants also used slightly less water. This, together with increased leaf and root cell osmotic potential, may reduce sensitivity to drought stress and improve plant growth [70]. Another study showed that root growth was stimulated in the presence of 0.75% biochar. This again facilitated water uptake and soil nutrient acquisition and, therefore, exerted beneficial effects on photosynthesis and lowered oxidative stress [74].

3.3. The Influence of Porosity on Soil Structure, Water and Gas Dynamics

Due to its porous nature, biochar can significantly increase gas transport in soil as well as a soil’s water-holding capacity (WHC) [7,75,76,77]. This is because pyrolysis results in an interconnected network of micropores, mesopores and macropores [41]. The distribution of pore sizes is governed by feedstock, with wood feedstock developing larger pores [75], and by pyrolysis conditions, with those biochars produced at high temperatures via a slow process more likely to produce more macropores (that is, greater than 50 μm in diameter) [75,78]. In addition, as biochar has approximately half the tensile strength of soil, it can reduce overall soil tensile strength, therefore reducing soil mechanical impedance [79], improving root elongation as well as mycorrhizal proliferation, thereby improving plant access to and assimilation of nutrients [7].
Biochar has been reported to improve crop yield through its effect on soil structure [80]. In some cases, biochar addition increased aggregate stability and reduced the detachment of colloidal material, improving soil structure [81]. However, in coarser soils, there was no such enhancement [82]. Where improvements were found, this may have been due to mechanisms such as carboxylic and phenolic functional groups on aged biochar surfaces, which form attachments with soil mineral surfaces. Also, a high CEC allows for cation bridge formation contributing to structural stabilisation [52].
Differences in biochar porosity resulting from different feedstocks have a direct influence on micro-organism population. This is because the adhesion of bacteria to biochar may be influenced by pore size [83]. Bacillus mucilaginosus and Acinetobacter sp. need a pore size of 2–4 μm if they are to enter [84]. In pores, they are better protected from dehydration and grazers and competitors. Surface tension holds water in the biochar, but it does so preferentially, with smaller pores exerting a greater holding capacity than larger ones. Equally, any increase in overall pore volume can increase water-holding capacity and provide greater resistance to water loss in drought-prone areas [85]. This balance is essential for nitrification, the optimum condition for which is 60% water-filled pore space (WFPS) [86]. The mode of application can have a critical influence, however. For instance, dry biochar is hydrophobic and may cause hydrophobicity in soil, interrupting water infiltration [87].
The effect of biochar on WFPS and soil aeration is often cited as the means through which complete denitrification is promoted and N2O emission reduced [21,88]. Denitrifiers are highly sensitive, requiring an oxygen concentration of less than 10% to denitrify [89]. The porous nature of biochar provides shelter, water and oxygen resulting in the rapid growth of heterotrophs and, therefore, the depletion of oxygen and the creation of anaerobic microsites [88]. Oxygen partial pressure strongly influences both denitrification and nitrification rates [89] and therefore any anaerobic microsites that may form around biochar particles may elevate N2O reduction activity [88].
Longer-term studies revealed the influence of different processes. For instance, in a seven-month study, biochar did not promote the reduction in N2O to N2, rather, the most prominent biochar-induced reduction in N2O resulted from an increase in metabolisable carbon [27]. Equally, in their work with fifteen different soil types, Cayuela et al. [90] found that the mechanism for reduction in N2O to N2 was not linked to an increase in soil aeration but was closely related to soil texture with fine soils promoting the last step of denitrification. Equally, where biochar would not significantly influence WFPS, they found that the effect of biochar on N2O production from denitrification did not correlate with the increased C:N ratio supplied by the biochar. Hence, micro-organism immobilisation of NO3 was not a driving mechanism for the observed N2O reductions [90]. Neither, it appears, was nitrite, which can have an inhibitory effect on the action of N2O reductase [91], reducing the production of the final product, N2. Hence, biochar, in reducing the release of a GHG, would be an appropriate amendment for agroecological systems, but again, the choice of biochar type, given the soil type, would need to be carefully considered if any increase in N2O was to be avoided or minimised.

3.4. The Sorption of Mineral Ions, Signalling Compounds, Heavy Metals and Organic Pollutants

Some biochars can be effective in adsorbing NH4+ and NO3 from the soil [92]. This apparent disadvantage, however, may lead to an increase in the ability of the soil ecosystem to feed plants as this reduced nitrogen availability to plant roots stimulates increased nodulation in legumes [64,93]. Root nodulation can influence the rate of biological fixation, and both the nodulation rate and development, as well as nitrogenase activity, can be affected by the presence of biochar. For instance, pyrolysised bamboo increased root nodulation by 243% and resulted in increased soybean growth [94]. However, another study found that three years after biochar application, regardless of the application rate, there was no significant difference in the total number of root nodules in clover between control and biochar-amended soil, although the level of nitrogenase activity in individual nodules was significantly higher in the biochar-treated soil [95].
Even though biochar can influence the structure of a micro-organism community to promote one nitrogen process or another, its sorption powers can often confound the result. In an experiment comparing biochar alone and biochar that had been shaken with dairy effluent for 24 h, both biochar treatments reduced net ammonification by 220% compared with soil alone. This suggested that the rate of nitrification was higher than the rate of ammonification. However, it appears that these rates were not changed in response to an increase in nitrifiers because CO2 emissions did not rise. Hence, it was postulated that the reduction in NH4+ was more likely due to its adsorption to biochar rather than immobilisation [96]. Similarly, ammonification was enhanced when a metabolisable organic nitrogen substrate was added to forest soil after fire, suggesting that the process is substrate-limited [97].
Again, biochar may influence the denitrification process by limiting micro-organism access to substrates. For instance, in one study, an acidic biochar absorbed NH4+, not only reducing the NH4+ leaching rate but also decreasing the NH3 volatilisation rate due to a reduction in substrate for the denitrification process [98]. However, the decrease in volatilisation rate was driven mainly by the acid–base reaction. Findings from another study concurred, concluding not only that the effectiveness of biochar to reduce N2O emission was, in part, dependent on the sorption of NH4+, decreasing the overall availability of nitrogen to denitrifiers, but that this ability changed over time [22]. With ageing, the effectiveness of biochar to reduce N2O emission (and NH4+ leaching) increased, as biochar surfaces become increasingly oxidised due to biotic and abiotic processes, potentially leading to an increase in cation exchange capacity [25], which may explain the reduction in available nitrogen [22]. Equally, biochar may also sorb N2O directly, thus reducing emission; however, sorption sites are likely to be taken up by water, carbon dioxide, organic matter and other mineral ions and the competition for these sites has not been elucidated, requiring further investigation [99].
The mechanism of root nodule formation and BNF in leguminous plants requires infection by nodule-forming bacteria. This process is governed by chemotaxis involving signalling pathways, which are initiated by polyphenolic signalling compounds (for example, flavonoids) released by the host plant [40,100,101,102,103,104]. Biochar is highly effective at adsorbing signalling compounds so any incorporation of biochar into soil, certainly at higher rates, may interfere with these signalling pathways, potentially interrupting nodule development and therefore nitrogen fixation [40,101,102,103,104].
However, biochar’s sorption capabilities may offer an advantage with regard to remediating contaminated soils [105,106,107]. Some organochlorine pesticides, agrochemicals, and other environmental contaminants induced, inhibited or delayed the recruitment of Rhizobia bacteria to host plant roots with the result that fewer root nodules are produced and lower rates of nitrogenase activity are seen [108]. However, the adsorption of these environmental pollutants to biochar has the effect of reducing toxicity to other soil micro-organisms, thereby increasing microbial biomass, including free-living nitrogen-fixing bacteria such as Bradyrhizobium japonicum [40,109].
It has been postulated that biochar can influence net nitrification rates through the sorption of inhibitory substances, but the mechanisms are site-specific and complex. For instance, biochar can sorb, and therefore reduce, the activity of compounds that could inhibit nitrifying bacteria [110] or potentially reduce the complexation of nitrogenous molecules, such as proteins, into tannin complexes [102]. In addition, over time, surface functional groups on aged biochar alter its capacity for absorption of different enzymes, thereby affecting enzyme activity and substrates, for instance, absorbing NH4+ and reducing nitrification [107,111]. The effect of the sorption capacity of biochar has not been elucidated for all soils but could profoundly influence nitrification rates in organic systems. Hence, this remains another area that requires further investigation to ensure that biochar’s different sorption mechanisms promote mineralisation processes such that crops are adequately supported.

3.5. The Influence of pH, Cation Exchange Capacity and Electron Shuttle Services

The pH of soil is critical because it affects plant nutrient availability by controlling the chemical forms of various nutrients and influencing the chemical reactions they undergo. For instance, phosphorus, molybdenum and calcium become increasingly unavailable with decreasing pH, with a corresponding decrease in crop productivity [112]. Biochar influences pH because it is generally alkaline due to its ash content and release of base cations, but also due to intrinsic alkaline organic functional groups [13]. However, the pH of a biochar is governed by both feedstock and pyrolysis conditions. Streubel et al. [14] found that the pH of herbaceous biochars was two units higher (9.4) than woody biochars (7.4) due to higher concentrations of ash in their study on contrasting biochar types (all pyrolysed at 350 °C). Yuan and Xu [15] found that pH was higher in biochars from leguminous feedstock with a pH range of 9.02 to 10.35 than in non-leguminous feedstock with a pH range of 8.00 to 9.24. Carbonates and organic anions of carboxyl and phenolic groups were the main alkalis but, again, this varied with feedstock.
As pH influences the chemical form and availability of substrates, it can affect change in both ammonia-oxidising archaea (AOA) and ammonia-oxidising bacteria (AOB) communities, thereby affecting mineralisation rates [113,114]. Accordingly, the addition of biochar has been shown to increase fixation rates, albeit to a lesser extent, and this capability appears to diminish over time [31,48].
pH, as amended by biochar, may have a greater influence on mineralisation rates, but the results are highly inconsistent and will depend on the existing pH levels of the amended soil. For instance, autotrophic nitrification generally occurs in neutral and alkaline soils because a critical enzyme, ammonia monooxygenase, uses NH3 as a substrate rather than NH4+ with the balance affected by pH, with a higher pH favouring NH3 [115]. Hence, although in an already alkaline soil, the addition of biochar resulted in a decrease in the number of nitrifiers [55]. However, an amended-acid soil resulted in a significant increase in the abundance of AOB correlating with an increased pH resulting from wheat biochar application [49]. Yet, the same study found no significant difference in the size of the AOA population with increasing pH, which correlates with the findings of [115]. This may be because AOA can be found in a wide range of soil pH, with some populations adapted to highly acidic soils. In another study, significantly lower net mineralisation rates have resulted from increased Eucalyptus biochar application due to decreased activity of the micro-organism community [116]. However, where there has been a decrease in nitrifiers, it has led to an increase in nitrification [55]. This increase may be because of biochar’s effect on the air and water balance in soil (discussed below).
Again, as with mineralisation, the pH effect on denitrification is inconsistent [90,117]. For instance, Borken et al. [118] found a decrease in N2O emission after liming of different forest soils, but Clough et al. (2004) found that WFPS had a greater bearing. However, Obia et al. [13] investigated the effect of two different types of biochar treated to remove alkalinity. They found that denitrification rates and gaseous products (NO, N2O and N2) were related to the increase in pH resulting from increased rates of biochar application. The untreated biochar suppressed NO and N2O, but increased N2 production, irrespective of the effect on denitrification rates. The treated biochar (which had been acid leached to reduce its liming effect) reduced or eliminated both its ability to suppress N2O and NO production, apparently confirming the importance of altered soil pH as a result of biochar addition for denitrification.
For comparison, in a study which increased the pH of soil with ash applications (as opposed to biochar), there were no observable reductions in N2O emissions [117]. The explanation proffered by Cayuela et al. [90] was the potential role of biochar as a reducing agent. This is because biochar may comprise manganese and iron—which readily function as electron acceptors [119]. Biochar may provide electron shuttle services, acting as an electrical conduit and facilitating electron transfer to micro-organisms. Hence, biochar would effectively compete with NO3 as an electron sink, thereby explaining a reduction in denitrification [90].
Biochar retains mineral ions in the rooting zone through its CEC. Feedstock type and pyrolysis conditions have an effect on the consequent negative surface charge of biochar with CEC including 3.8, 60.6, 137.6 and 254 cmol kg−1 for sugarcane bagasse, rice straw, chicken manure and peanut straw, respectively [120,121]. This surface charge results from carboxylate groups on the surface of biochar itself but also from exposed carboxylate groups of organic acids sorbed onto the biochar [122]. However, Wu et al. [123] found that, in the case of rice straw biochar, it was pyrolysis temperature, rather than residence time, which had a greater bearing on CEC. Higher pyrolysis temperatures generally cause greater condensation of aromatic structures resulting in aromatic carbon forms with less surface area and fewer oxidisable functional groups [124]. Therefore, not all biochars can raise soil CEC or oxidise to do so over time [120]. In fact, although many studies report an increase in CEC of soils amended with biochar [12], these are often degraded, poor soils with an inherently low CEC. For example, work done by Martinsen et al. [125] reported the influence of three different biochars on 31 different soils, which were all acidic and had a low- to medium-range CEC. Here, the addition of biochar was found to raise CEC as well as pH and exchangeable bases. However, where CEC increases over time, it may result in increased retention of NH4+ and, therefore, limit the supply of this substrate for other processes, causing a decrease in BNF [126,127].
Biochar CEC can have a direct effect on NH4+ leaching. Singh et al. [32] found that, over time, soils amended with biochar became effective in reducing NH4+ leaching but efficacy varied with pyrolysis temperature, with the high-temperature biochars decreasing NH4+ leaching from both soils (Alfisol and Vertisol) by 55–93%, but low-temperature biochars decreased leaching by 87–94% in the Vertisol only. This may have been due to the reduced surface area in the biochar resulting from low-temperature pyrolysis [71,75].
Different biochars can have some level of anion exchange capacity (AEC). Maize stover, cellulose, alfalfa meal and albumin biochars ranged from 0.602 to 27.76 cmol kg−1, respectively, and this increases with increasing pH [128]. These results may explain the reduction in NO3 leaching from biochar-amended soil treated with a biosolid where leaching of NO3 decreased to a level below control treatments [129].

3.6. The Influence of Inhibitory Substances

After pyrolysis of biomass, compounds toxic to micro-organisms may be present including polyaromatic hydrocarbons (PAHs) [130,131,132]. The presence of inhibitory substances due to biochar addition is of particular concern as once added to soil, it is almost impossible to remove, therefore any negative environmental consequences may be long-lasting. For instance, Anderson et al. [39] found that applying biochar to a silt-loam soil decreased the abundance of Nitrosovibrio—an AOB. As this is a rate-limiting step for nitrification, rates fell, which may be due to the introduction of inhibitory substances [33]. Wang et al. also found that phenolic compounds, which are retained by biochars, especially at low-pyrolysis temperatures, may inhibit microbial activity [133]. Furthermore, in their experiment with peanut shell biochar, one treatment with retained phenolic compounds and another without, they found that the presence of phenolic compounds likely reduced AOB abundance, thereby suppressing nitrification processes.

4. Summary of the Influence of Biochar on the Nitrogen Cycle

To conclude the above discussion, as the nitrogen cycle is governed by the activities of micro-organisms and biochar influences both micro-organism action and community composition, as well as other physical and chemical processes in the soil, it has a profound influence on fixation, assimilation, mineralisation and denitrification.
The main influences on fixation are through the promotion of mycorrhizal root colonisation due to the provision of carbohydrates that form part of the metabolisable fraction of carbon supplied by biochar or by providing physical protection from fungal grazers [48,68,69]. Biochar sorption capabilities may be hugely influential as they sorb nitrogen, reducing its availability in soil, thereby promoting nodulation [134]. It sorbs pollutants detrimental to micro-organism growth including fixers [40,109], but can also interfere with signal pathways, potentially interrupting nodule development [101,103]. Biochar can alter soil pH such that soil fixers benefit, although this process may diminish over time [135]. The provision of a metabolisable form of carbon, as well as recalcitrant forms, has been found to increase nitrogen-fixing organisms [31,48]. Figure 2 gives a summary of the substrates and products of fixation and how biochar influences the chemical and biological processes that lead to the products.
Biochar influences the assimilation of nitrogen into micro-organisms and plants via several mechanisms. Firstly, it can reduce leaching, keeping nitrogen in the rhizosphere and available for uptake. It can do this through the adsorption of NH4 or organic nitrogen onto biochar, intercalation or cation or anion exchange reactions [136]. Secondly, it improves soil water-holding capacity in some soils, which aids root and hyphal elongation and nutrient capture [68,135] even under conditions of high water evaporation stress [77]. Any change in pH brought about by biochar can either improve nutrient availability (and therefore leaching) or nutrients can become further unavailable [137]. In addition, the metabolisable carbon fraction can result in increased micro-organism growth and immobilisation of nitrogen [18,120,138,139]. These factors (captured in Figure 3) combine to create a complex soil ecosystem resulting in different rates of assimilation but also, therefore, different autotroph communities able to mineralise organic matter and continue the cycle.
Much research has been conducted on the influence of biochar on mineralisation rates and responses vary with soil type. However, the main mechanisms of influence include the provision of a metabolisable carbon, which influences micro-organism growth (Figure 4). For instance, as the process of ammonification is mediated by micro-organisms (usually Bacillus spp., Proteus spp. and Pseudomonas spp.) reliant on nutrient, energy and water resources, as well as communication mechanisms, and biochar can influence the rate and result of ammonification through its influence on soil structure and WHC, the provision and sorption of toxins, signalling compounds, nutrients and energy sources (that is, metabolisable carbon). However, studies analysing the effect of biochar on the transformation of NH3 or NH4+ to NO3 or nitrification have revealed contrasting results. In some soils, there is no effect on nitrification rates [56,110]. However, in others, biochar promoted net nitrification [139,140].
If nitrogen is limited in that soil, this could lead to the mineralisation of recalcitrant organic sources of nitrogen [28,39]. Any change in pH can affect the abundance and diversity of mineralising micro-organisms [39]. Sorption of substrates (NH4+) can limit mineralisation [136], but sorption of pollutants may benefit nitrifying populations [40]. Biochar pores may offer protection from predation and dehydration, thereby protecting mineralising populations [48,69].

The Influence of Biochar on Denitrification

The denitrification process is generally governed by several species of heterotrophic facultative anaerobic bacteria and archaea, which oxidise NH3 and NH4+ via more than one enzymatic pathway [141]. The composition and genetic capability of the denitrifying micro-organism community and environmental conditions of a given soil dictate the pathways and partitioning of eventual nitrogen products. However, the facultative nature of the anaerobic bacteria, enzyme action and enzymatic pathway and the reduction in oxidised forms of nitrogen in response to an electron donor governs the mode of action biochar has on these processes. Hence, the mechanisms of the effect of biochar on denitrification known so far include the provision of metabolisable carbon and subsequent effect on soil aeration and the development of anaerobic microsites; the adsorption of substrates and products, for example, N2O; pH; and the provision of electron shuttle services and toxic compounds (Figure 5).

5. Conclusions

This review has assessed biochar along different parameters in terms of their influence on soil nitrogen transformations. Biochar’s influence on fixation rates generally stems from its provision of two forms of carbon, metabolisable and recalcitrant, both of which have been found to impact the number of nitrogen-fixing organisms. Carbon forms also affect assimilation rates as a high metabolisable carbon fraction can result in nitrogen immobilisation, and incidences of a decrease in yield have been noted. This clearly has implications for the sustainable use of biochar in future, where agricultural yield is a consideration. However, biochar also changes pH, making nutrients more or less available for assimilation as well as affecting abundance and diversity of mineralising micro-organisms. Equally, biochar’s ability to sorb pollutants may benefit nitrifying populations. Biochar also influences soil aeration and development of anaerobic microsites thereby influencing the release of denitrification products. However, the rates of such processes are dependent on biochar and soil type. Future studies should ensure that a detailed analysis of biochar feedstock, pyrolysis conditions and biochar characteristics are included, which would enable a greater understanding of the role of biochar on nitrogen transformations and should ensure sustainable deployment of this proven climate change mitigation tool. It is hoped that this review will serve as guidance for future studies and as a reference for students and practitioners alike.

Author Contributions

All authors contributed to the study conception and design. Material preparation, data collection and review were performed by D.D. The first draft of the manuscript was written by D.D., F.R., S.C. and R.L. commented on all versions of the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kaza, S.; Yao, L.C.; Bhada-Tata, P.; Van Woerden, F. What a Waste 2.0: A Global Snapshot of Solid Waste Management to 2050; Urban Development Series; World Bank: Washington, DC, USA, 2018. [Google Scholar]
  2. IEA. Global Energy Review 2020; IEA: Paris, France, 2020. [Google Scholar]
  3. Rickson, R.J.; Deeks, L.K.; Graves, A.; Harris, J.A.H.; Kibblewhite, M.G.; Sakrabani, R. Input constraints to food production: The impact of soil degradation. Food Secur. 2015, 7, 351–364. [Google Scholar] [CrossRef]
  4. Dekker, S.C.; Kraneveld, A.D.; van Dijk, J.; Kalfagianni, A.; Knulst, A.C.; Lelieveldt, H.; Moors, E.H.M.; Müller, E.; Pieters, R.H.H.; Pieterse, C.M.J.; et al. Towards Healthy Planet Diets—A Transdisciplinary Approach to Food Sustainability Challenges. Challenges 2020, 11, 21. [Google Scholar] [CrossRef]
  5. Godfray, H.C.J.; Crute, I.R.; Haddad, L.; Lawrence, D.; Muir, J.F.; Nisbett, N.; Pretty, J.; Robinson, S.; Toulmin, C.; Whiteley, R. The future of the global food system. Philos. Trans. R. Soc. B Biol. Sci. 2010, 365, 2769–2777. [Google Scholar] [CrossRef] [PubMed]
  6. McNaught, A.D.; Wilkinson, A. IUPAC. Compendium of Chemical Terminology (the Gold Book), 2nd ed.; Blackwell Scientific Publications: Oxford, UK, 1997. [Google Scholar]
  7. 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]
  8. Seifritz, W. Should we store carbon in charcoal? Int. J. Hydrogen Energy 1993, 18, 405–407. [Google Scholar] [CrossRef]
  9. Sombroek, W.G.; Nachtergaele, F.O.; Hebel, A. Amounts, dynamics and sequestering of carbon in tropical and subtropical soils. Ambio 1993, 22, 417–426. [Google Scholar]
  10. 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]
  11. Ding, Y.; Liu, Y.-X.; Wu, W.-X.; Shi, D.-Z.; Yang, M.; Zhong, Z.-K. Evaluation of Biochar Effects on Nitrogen Retention and Leaching in Multi-Layered Soil Columns. Water Air Soil Pollut. 2010, 213, 47–55. [Google Scholar] [CrossRef]
  12. Tan, S.; Zhou, G.; Yang, Q.; Ge, S.; Liu, J.; Cheng, Y.W.; Yek, P.N.Y.; Mahari, W.A.W.; Kong, S.H.; Chang, J.-S.; et al. Utilization of current pyrolysis technology to convert biomass and manure waste into biochar for soil remediation: A review. Sci. Total Environ. 2023, 864, 160990. [Google Scholar] [CrossRef]
  13. Obia, A.; Cornelissen, G.; Mulder, J.; Dörsch, P. Effect of Soil pH Increase by Biochar on NO, N2O and N2 Production during Denitrification in Acid Soils. PLoS ONE 2015, 10, e0138781. [Google Scholar] [CrossRef]
  14. Streubel, J.D.; Collins, H.P.; Garcia-Perez, M.; Tarara, J.; Granatstein, D.; Kruger, C. Influence of Contrasting Biochar Types on Five Soils at Increasing Rates of Application. Soil Sci. Soc. Am. J. 2011, 75, 1402–1413. [Google Scholar] [CrossRef]
  15. Yuan, J.; Xu, R. Effects of biochars generated from crop residues on chemical properties of acid soils from tropical and subtropical China. Soil Res. 2012, 50, 570–578. [Google Scholar] [CrossRef]
  16. Lehmann, J.; Pereira da Silva, J.; Steiner, C.; Nehls, T.; Zech, W.; Glaser, B. Nutrient availability and leaching in an archaeological Anthrosol and a Ferralsol of the Central Amazon basin: Fertilizer, manure and charcoal amendments. Plant Soil 2003, 249, 343–357. [Google Scholar] [CrossRef]
  17. Marchetti, R.; Castelli, F.; Orsi, A.; Sghedoni, L.; Bochicchio, D. Biochar from swine manure solids: Influence on carbon sequestration and Olsen phosphorus and mineral nitrogen dynamics in soil with and without digestate incorporation. Ital. J. Agron. 2012, 7, e26. [Google Scholar] [CrossRef]
  18. Deenik, J.L.; Cooney, M.J. The Potential Benefits and Limitations of Corn Cob and Sewage Sludge Biochars in an Infertile Oxisol. Sustainability 2016, 8, 131. [Google Scholar] [CrossRef]
  19. Kelley, K.R.; Stevenson, F.J. ;Characterization and extract ability of immobilized 15N from the soil microbial biomass. Soil Biol. Biochem. 1985, 17, 517–523. [Google Scholar] [CrossRef]
  20. Haider, G.; Steffens, D.; Moser, G.; Müller, C.; Kammann, C.I. Biochar reduced nitrate leaching and improved soil moisture content without yield improvements in a four-year field study. Agric. Ecosyst. Environ. 2017, 237, 80–94. [Google Scholar] [CrossRef]
  21. Bateman, E.; Baggs, E.M. Contributions of nitrification and denitrifcation to N2O emissions from soils at different water-filled pore space. Biol. Fertil. Soils Coop. J. Int. Soc. Soil Sci. 2005, 41, 379–388. [Google Scholar]
  22. Singh, B.; Singh, B.P.; Cowie, A. Characterisation and evaluation of biochars for their application as a soil amendment. Aust. J. Soil Res. 2010, 48, 516. [Google Scholar] [CrossRef]
  23. Kimetu, J.; Lehmann, J. Stability and stabilisation of biochar and green manure in soil with different organic carbon contents. Soil Res. 2010, 48, 577–585. [Google Scholar] [CrossRef]
  24. Yeboah, E.; Ofori, P.; Quansah, G.W.; Dugan, E.; Sohi, S.P. Improving soil productivity through biochar amendments to soils. Afr. J. Environ. Sci. Technol. 2009, 3, 34. [Google Scholar]
  25. Cheng, C.; Lehmann, J. Ageing of black carbon along a temperature gradient. Chemosphere 2009, 75, 1021–1027. [Google Scholar] [CrossRef] [PubMed]
  26. 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]
  27. Ameloot, N.; Maenhout, P.; De Neve, S.; Sleutel, S. Biochar-induced N2O emission reductions afer field incorporation in a loam soil. Geoderma 2016, 267, 10–16. [Google Scholar] [CrossRef]
  28. Nelissen, V.; Rütting, T.; Huygens, D.; Staelens, J.; Ruysschaert, G.; Boeckx, P. Maize biochars accelerate short-term soil nitrogen dynamics in a loamy sand soil. Soil Biol. Biochem. 2012, 55, 20–27. [Google Scholar] [CrossRef]
  29. Rittl, T.F.; Oliveira, D.M.S.; Canisares, L.P.; Sagrilo, E.; Butterbach-Bahl, K.; Dannenmann, M.; Cerri, C.E.P. High Application Rates of Biochar to Mitigate N2O Emissions from a N-Fertilized Tropical Soil under Warming Conditions. Front. Environ. Sci. 2021, 8, 1. [Google Scholar] [CrossRef]
  30. Lee, S.-I.; Park, H.-J.; Jeong, Y.-J.; Seo, B.-S.; Kwak, J.-H.; Yang, H.I.; Xu, X.; Tang, S.; Cheng, W.; Lim, S.-S.; et al. Biochar-induced reduction of N2O emission from East Asian soils under aerobic conditions: Review and data analysis. Environ. Pollut. 2021, 291, 118154. [Google Scholar] [CrossRef] [PubMed]
  31. 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. [Google Scholar]
  32. Singh, B.P.; Hatton, B.J.; Singh, B.; Cowie, A.L.; Kathuria, A. Influence of Biochars on Nitrous Oxide Emission and Nitrogen Leaching from Two Contrasting Soils. J. Environ. Qual. 2010, 39, 1224–1235. [Google Scholar] [CrossRef]
  33. Clough, T.J.; Condron, L.M. Biochar and the nitrogen cycle. J. Environ. Qual. 2010, 39, 1218–1223. [Google Scholar] [CrossRef]
  34. IPCC. Climate Change 2007: Synthesis Report; Intergovernmental Panel on Climate Change (IPCC): Geneva, Switzerland, 2007. [Google Scholar]
  35. Vitousek, P.M.; Menge, D.N.; Reed, S.C.; Cleveland, C.C. Biological nitrogen fixation: Rates, patterns and ecological controls in terrestrial ecosystems. Philos. Trans. R. Soc. B Biol. Sci. 2013, 368, 1621. [Google Scholar] [CrossRef] [PubMed]
  36. Canali, S.; Di Bartolomeo, E.; Tittarelli, F.; Montemurro, F.; Verrastro, V.; Ferri, D. Comparison of different laboratory incubation procedures to evaluate nitrogen mineralization in soils amended with aerobic and anaerobic stabilized organic materials. J. Food Agric. Environ. 2011, 9, 540–546. [Google Scholar]
  37. Strock, J.S. Ammonification. In Encyclopedia of Ecology; Elsevier B.V.: Amsterdam, The Netherlands, 2008; pp. 162–165. [Google Scholar]
  38. Crutzen, P.J. The influence of nitrogen oxides on the atmospheric ozone content. QJR Meteorol. Soc. 1970, 96, 320–325. [Google Scholar] [CrossRef]
  39. 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]
  40. Seneviratne, M.; Weerasundara, L.; Ok, Y.S.; Rinklebe, J.; Vithanage, M. Phytotoxicity attenuation in Vigna radiata under heavy metal stress at the presence of biochar and N fixing bacteria. J. Environ. Manag. 2017, 186, 293–300. [Google Scholar] [CrossRef] [PubMed]
  41. Shackley, S.; Ruysschaert, G.; Zwart, K.; Glaser, B. (Eds.) Biochar in European Soils and Agriculture: Science and Practice; Earthscan from Routledge: London, UK, 2016. [Google Scholar]
  42. Fawzy, S.; Osman, A.I.; Yang, H.; Doran, J.; Rooney, D.W. Industrial biochar systems for atmospheric carbon removal: A review. Environ. Chem. Lett. 2021, 19, 3023. [Google Scholar] [CrossRef]
  43. Thies, J.; Rillig, M. Characteristics of biochar: Biological properties. In Biochar for Environmental Management; Lehmann, J., Joseph, S., Eds.; Earthscan: Oxford, UK, 2009; p. 85. [Google Scholar]
  44. Shackley, S.; Sohi, S.; Brownsort, P.; Carter, S.; Cook, J.; Cunningham, C.; Gaunt, J.; Hammond, J.; Ibarrola, R.; Mašek, O.; et al. An Assessment of the Benefits and Issues Associated with the Application of Biochar to Soil; Department for Environment, Food and Rural Affairs: London, UK, 2010.
  45. 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]
  46. Bruun, E.W.; Müller-Stöver, D.; Ambus, P.; Hauggaard-Nielsen, H. Application of biochar to soil and N2O emissions: Potential effects of blending fast-pyrolysis biochar with anaerobically digested slurry. Eur. J. Soil Sci. 2011, 62, 581–589. [Google Scholar] [CrossRef]
  47. Hagemann, N.; Harter, J.; Behrens, S. Elucidating the impacts of biochar applications on nitrogen cycling microbial communities. In Biochar Application; Ralebitso-Senior, T.K., Orr, C.H., Eds.; Elsevier Inc.: Amsterdam, The Netherlands, 2016; Chapter 7. [Google Scholar] [CrossRef]
  48. Güereña, D.T.; Lehmann, J.; Thies, J.E.; Enders, A.; Karanja, N.; Neufeldt, H. Partitioning the contributions of biochar properties to enhanced biological nitrogen fixation in common bean (Phaseolus vulgaris). Biol. Fertil. Soils 2015, 51, 479–491. [Google Scholar] [CrossRef]
  49. Zhang, Q.-Z.; Dijkstra, F.A.; Liu, X.-R.; Wang, Y.-D.; Huang, J.; Lu, N. Effects of Biochar on Soil Microbial Biomass after Four Years of Consecutive Application in the North China Plain. PLoS ONE 2014, 9, e102062. [Google Scholar] [CrossRef]
  50. Nguyen, B.T.; Lehmann, J.; Kinyangi, J.; Smernik, R.; Riha, S.J.; Engelhard, M.H. Long-term black carbon dynamics in cultivated soil. Biogeochemistry 2009, 92, 163–176. [Google Scholar] [CrossRef]
  51. Bird, M.I.; Moyo, C.; Veenendaal, E.M.; Lloyd, J.; Frost, P. Stability of elemental carbon in a savanna soil. Glob. Biogeochem. Cycles 1999, 13, 923–932. [Google Scholar] [CrossRef]
  52. Lin, Y.; Munroe, P.; Joseph, S.; Kimber, S.; Van Zwieten, L. Nanoscale organo-mineral reactions of biochars in ferrosol: An investigation using microscopy. Plant Soil 2012, 357, 369–380. [Google Scholar] [CrossRef]
  53. Fontaine, S.; Mariotti, A.; Abbadie, L. The priming effect of organic matter: A question of microbial competition? Soil Biol. Biochem. 2003, 35, 837–843. [Google Scholar] [CrossRef]
  54. Blagodatskaya, E.V.; Blagodatsky, S.A.; Anderson, T.; Kuzyakov, Y. Contrasting effects of glucose, living roots and maize straw on microbial growth kinetics and substrate availability in soil. Eur. J. Soil Sci. 2009, 60, 186–197. [Google Scholar] [CrossRef]
  55. Prommer, J.; Wanek, W.; Hofhansl, F.; Trojan, D.; Offre, P.; Urich, T.; Schleper, C.; Sassmann, S.; Kitzler, B.; Soja, G.; et al. Biochar Decelerates Soil Organic Nitrogen Cycling but Stimulates Soil Nitrification in a Temperate Arable Field Trial. PLoS ONE 2014, 9, e86388. [Google Scholar] [CrossRef] [PubMed]
  56. Cheng, Y.; Cai, Z.; Chang, S.X. Wheat straw and its biochar have contrasting effects on inorganic N retention and N2O production in a cultivated Black Chernozem. Biol. Fertil. Soils Coop. J. Int. Soc. Soil Sci. 2012, 48, 941–946. [Google Scholar] [CrossRef]
  57. Wang, X.; Zhou, W.; Liang, G.; Song, D.; Zhang, X. Characteristics of maize biochar with different pyrolysis temperatures and its effects on organic carbon, nitrogen and enzymatic activities after addition to fluvo-aquic soil. Sci. Total Environ. 2015, 538, 137–144. [Google Scholar] [CrossRef]
  58. García-Sánchez, M.; Šípková, A.; Száková, J.; Kaplan, L.; Ochecová, P.; Tlustoš, P. Applications of organic and inorganic amendments induce changes in the mobility of mercury and macro- and micronutrients of soils. Sci. World J. 2014, 2014, 407049. [Google Scholar] [CrossRef]
  59. Huang, Y.; Zou, J.; Zheng, X.; Wang, Y.; Xu, X. Nitrous oxide emissions as influenced by amendment of plant residues with different C:N ratios. Soil Biol. Biochem. 2004, 36, 973–981. [Google Scholar] [CrossRef]
  60. Robertson, G.E.; Groffman, P.M. Nitrogen transformations. In Soil Microbiology and Biochemistry; Paul, E.A., Ed.; Elsevier Academic Press: Oxford, UK, 2007; pp. 341–362. [Google Scholar]
  61. Millar, N.; Baggs, E.M. Relationships between N2O emissions and water-soluble C and N contents of agroforestry residues after their addition to soil. Soil Biol. Biochem. 2005, 37, 605–608. [Google Scholar] [CrossRef]
  62. Mosa, A.; Mansour, M.M.; Soliman, E.; El-Ghamry, A.; El Alfy, M.; El Kenawy, A.M. Biochar as a Soil Amendment for Restraining Greenhouse Gases Emission and Improving Soil Carbon Sink: Current Situation and Ways Forward. Sustainability 2023, 15, 1206. [Google Scholar] [CrossRef]
  63. Udall, D.; Rayns, F.; Mansfield, T. LIVING SOILS: A Call to Action; Soil Association: Bristol, UK.; Centre for Agroecology, Water and Resilience (CAWR) at Coventry University: Coventry, UK, 2014. [Google Scholar]
  64. Lehmann, J.; Joseph, S. Biochar for Environmental Management: Science and Technology, 1st ed.; Earthscan: London, UK, 2009. [Google Scholar]
  65. Awad, Y.M.; Blagodatskaya, E.; Ok, Y.S.; Kuzyakov, Y. Effects of polyacrylamide, biopolymer, and biochar on decomposition of soil organic matter and plant residues as determined by 14C and enzyme activities. Eur. J. Soil Biol. 2012, 48, 1–10. [Google Scholar] [CrossRef]
  66. Mukome, F.N.D.; Zhang, X.; Silva, L.C.; Six, J.; Parikh, S.J. Use of chemical and physical characteristics to investigate trends in biochar feedstocks. J. Agric. Food Chem. 2013, 61, 2196–2204. [Google Scholar] [CrossRef] [PubMed]
  67. Scheifele, M.; Hobi, A.; Buegger, F.; Gattinger, A.; Schulin, R.; Boller, T.; Mäder, P. Impact of pyrochar and hydrochar on soybean (Glycine max L.) root nodulation and biological nitrogen fixation. J. Plant Nutr. Soil Sci. 2017, 180, 199–211. [Google Scholar]
  68. Warnock, D.D.; Lehmann, J.; Kuyper, T.W.; Rillig, M.C. Mycorrhizal responses to biochar in soil—Concepts and mechanisms. Plant Soil 2007, 300, 9–20. [Google Scholar] [CrossRef]
  69. Vanek, S.J.; Lehmann, J. Phosphorus availability to beans via interactions between mycorrhizas and biochar. Plant Soil 2015, 395, 105. [Google Scholar] [CrossRef]
  70. 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]
  71. Major, J.; Lehmann, J.; Rondon, 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]
  72. González, J.A.; Gallardo, M.; Hilal, M.B.; Rosa, M.D.; Prado, F.E. Physiological responses of quinoa (Chenopodium quinoa Willd.) to drought and water logging stresses; dry matter partitioning. Bot. Stud. 2009, 50, 35–42. [Google Scholar]
  73. Domingues, R.R.; Trugilho, P.F.; Silva, C.A.; Melo, I.C.N.D.; Melo, L.C.; Magriotis, Z.M.; Sanchez-Monedero, M.A. Properties of biochar derived from wood and high-nutrient biomasses with the aim of agronomic and environmental benefits. PLoS ONE 2017, 12, e0176884. [Google Scholar] [CrossRef] [PubMed]
  74. Abideen, Z.; Koyro, H.; Huchzermeyer, B.; Ansari, R.; Zulfiqar, F.; Gul, B. Ameliorating effects of biochar on photosynthetic efficiency and antioxidant defence of Phragmites karka under drought stress. Plant Biol. J. 2020, 22, 259–266. [Google Scholar] [CrossRef] [PubMed]
  75. Downie, A.; Crosky, A.; Munroe, P. Physical properties of biochar. In Biochar for Environmental Management; Lehmann, J., Joseph, S., Eds.; Routledge: Abingdon, UK, 2009; Chapter 2. [Google Scholar]
  76. Glaser, B.; Lehmann, J.; Zech, W. Ameliorating physical and chemical properties of highly weathered soils in the tropics with charcoal—A review. Biol. Fertil. Soils 2002, 35, 219–230. [Google Scholar] [CrossRef]
  77. Ghorbani, M.; Neugschwandtner, R.W.; Konvalina, P.; Asadi, H.; Kopecký, M.; Amirahmadi, E. Comparative effects of biochar and compost applications on water holding capacity and crop yield of rice under evaporation stress: A two-years field study. Paddy Water Environ. 2023, 21, 47–58. [Google Scholar] [CrossRef]
  78. Guizani, C.; Jeguirim, M.; Valin, S.; Limousy, L.; Salvador, S. Biomass Chars: The Effects of Pyrolysis Conditions on Their Morphology, Structure, Chemical Properties and Reactivity. Energies 2017, 10, 796. [Google Scholar] [CrossRef]
  79. Atkinson, C.J.; Fitzgerald, J.D.; Hipps, N.A. Potential mechanisms for achieving agricultural benefits from biochar application to temperature soils: A review. Plant Soil 2010, 337, 1–18. [Google Scholar] [CrossRef]
  80. Liu, Z.; Chen, X.; Jing, Y.; Li, Q.; Zhang, J.; Huang, Q. Effects of biochar amendment on rapeseed and sweet potato yields and water stable aggregate in upland red soil. Catena 2014, 123, 45–51. [Google Scholar] [CrossRef]
  81. Soinne, H.; Hovi, J.; Tammeorg, P.; Turtola, E. Effect of biochar on phosphorus sorption and clay soil aggregate stability. Geoderma 2014, 219–220, 162–167. [Google Scholar] [CrossRef]
  82. Wang, Y.; Hu, N.; Ge, T.; Kuzyakov, Y.; Wang, Z.-L.; Li, Z.; Tang, Z.; Chen, Y.; Wu, C.; Lou, Y. Soil aggregation regulates distributions of carbon, microbial community and enzyme activities after 23-year manure amendment. Appl. Soil Ecol. 2017, 111, 65–72. [Google Scholar] [CrossRef]
  83. Rivera-Utrilla, J.; Bautista-Toledo, I.; Ferro-García, M.A.; Moreno-Castilla, C. Activated carbon surface modifications by adsorption of bacteria and their effect on aqueous lead adsorption. J. Chem. Technol. Biotechnol. 2001, 76, 1209–1215. [Google Scholar] [CrossRef]
  84. Samonin, V.V.; Elikova, E.E. A study of the adsorption of bacterial cells on porous materials. Microbiology 2004, 73, 696–701. [Google Scholar] [CrossRef]
  85. Hueso, S.; Hernández, T.; García, C. Resistance and resilience of the soil microbial biomass to severe drought in semiarid soils: The importance of organic amendments. Appl. Soil Ecol. 2011, 50, 27–36. [Google Scholar] [CrossRef]
  86. Linn, D.M.; Doran, J.W. Effect of Water-filled Pore-space on Carbon-dioxide and Nitrous-oxide Production in Tilled and Nontilled Soils. Soil Sci. Soc. Am. J. 1984, 48, 1267–1272. [Google Scholar] [CrossRef]
  87. Mao, J.; Zhang, K.; Chen, B. Linking hydrophobicity of biochar to the water repellency and water holding capacity of biochar-amended soil. Environ. Pollut. 2019, 253, 779–789. [Google Scholar] [CrossRef]
  88. van Zwieten, L.; Singh, B.P.; Joseph, S.; Kimber, S.; Cowie, A.; Chan, Y. Biochar and the emissions of non-CO2 greenhouse gases from soil. In Biochar for Environmental Management, Science and Technology; Lehmann, J., Joseph, S., Eds.; Earthscan: London, UK, 2009. [Google Scholar]
  89. Khalil, K.; Mary, B.; Renault, P. Nitrous oxide production by nitrification and denitrification in soil aggregates as affected by O2 concentration. Soil Biol. Biochem. 2004, 36, 687–699. [Google Scholar] [CrossRef]
  90. Cayuela, M.L.; Sánchez-Monedero, M.A.; Roig, A.; Hanley, K.; Enders, A.; Lehmann, J. Biochar and denitrification in soils: When, how much and why does biochar reduce N2O emissions? Sci. Rep. 2013, 3, 1732. [Google Scholar] [CrossRef]
  91. Firestone, M.K.; Smith, M.S.; Firestone, R.B.; Tiedje, J.M. Influence of nitrate, nitrite, and oxygen on the composition of the gaseous products of denitrification in soil. Soil Sci. Soc. Am. J. 1979, 43, 1140–1144. [Google Scholar] [CrossRef]
  92. Rivka, F.B.; Laird, D.A.; Spokas, K.A. Sorption of ammonium and nitrate to biochars is electrostatic and pH-dependent. Sci. Rep. 2018, 8, 17627. [Google Scholar]
  93. Gao, S.; Thomas, H.D.L. Influence of biochar on soil nutrient transformations, nutrient leaching, and crop yield. Adv. Plants Agric. Res. 2016, 4, 348–362. [Google Scholar]
  94. Wang, C.; Alidoust, D.; Yang, X.; Isoda, A. Effects of bamboo biochar on soybean root nodulation in multi-elements contaminated soils. Ecotoxicol. Environ. Saf. 2018, 150, 62–69. [Google Scholar] [CrossRef]
  95. Quilliam, R.S.; DeLuca, T.H.; Jones, D.L. Biochar application reduces nodulation but increases nitrogenase activity in clover. Plant Soil 2013, 366, 83. [Google Scholar] [CrossRef]
  96. Sarkhot, D.V.; Berhe, A.A.; Ghezzehei, T.A. Impact of biochar enriched with dairy manure effluent on carbon and nitrogen dynamics. J. Environ. Qual. 2012, 41, 1107–1114. [Google Scholar] [CrossRef] [PubMed]
  97. DeLuca, T.H.; Nilsson, M.C.; Zackrisson, O. Nitrogen mineralization and phenol accumulation along a fire chronosequence in northern Sweden. Oecologia 2002, 133, 206–214. [Google Scholar] [CrossRef] [PubMed]
  98. Esfandbod, M.; Phillips, I.R.; Miller, B.; Rashti, M.R.; Lan, Z.M.; Srivastava, P.; Singh, B.; Chen, C.R. Aged acidic biochar increases nitrogen retention and decreases ammonia volatilisation in alkaline bauxite residue sand. Ecol. Eng. 2017, 98, 157–165. [Google Scholar] [CrossRef]
  99. Cornelissen, G.; Rutherford, D.W.; Arp, H.P.H.; Dörsch, P.; Kelly, C.N.; Rostad, C.E.; Cornelissen, G.; Rutherford, D.W.; Arp, H.P.H.; Dörsch, P. Sorption of pure N2O to biochars and other organic and inorganic materials under anhydrous conditions. Environ. Sci. Technol. 2013, 47, 7704–7712. [Google Scholar] [CrossRef]
  100. Slattery, J.F.; Coventry, D.R.; Slattery, W.J. Rhizobial ecology as affected by the soil environment. Aust. J. Exp. Agric. 2001, 41, 289–298. [Google Scholar] [CrossRef]
  101. Beare, M.H.; Gregorich, E.G.; St-Georges, P. Compaction effects on CO2 and N2O production during drying and rewetting of soil. Soil Biol. Biochem. 2009, 41, 611–621. [Google Scholar] [CrossRef]
  102. DeLuca, T.H.; Gundale, M.J.; MacKenzie, M.D.; Jones, D.L. Biochar effects on soil nutrient transformations. In Biochar for Environmental Management; Lehmann, J., Joseph, S., Eds.; Earthscan: London, UK, 2010; p. 419. [Google Scholar]
  103. Gundale, M.J.; Nilsson, M.C.; Pluchon, N.; Wardle, D.A. The Effect of Biochar Management on Soil and Plant Community Properties in a Boreal Forest. GCB Bioenergy 2016, 8, 777–789. [Google Scholar] [CrossRef]
  104. Ni, J.; Pignatello, J.; Xing, B. Adsorption of Aromatic Carboxylate Ions to Black Carbon (Biochar) Is Accompanied by Proton Exchange with Water. Environ. Sci. Technol. 2011, 45, 9240–9248. [Google Scholar] [CrossRef]
  105. Tang, J.; Zhu, W.; Kookana, R.; Katayama, A. Characteristics of biochar and its application in remediation of contaminated soil. J. Biosci. Bioeng. 2013, 116, 653–659. [Google Scholar] [CrossRef]
  106. Zhang, X.; Wang, H.; He, L.; Lu, K.; Sarmah, A.; Li, J.; Bolan, N.S.; Pei, J.; Huang, H. Using biochar for remediation of soils contaminated with heavy metals and organic pollutants. Environ. Sci. Pollut. Res. Int. 2013, 20, 8472–8483. [Google Scholar] [CrossRef] [PubMed]
  107. Zhu, X.; Chen, B.; Zhu, L.; Xing, B. Effects and mechanisms of biochar-microbe interactions in soil improvement and pollution remediation: A review. Environ. Pollut. 2017, 227, 98–115. [Google Scholar] [CrossRef] [PubMed]
  108. Fox, J.E.; Gulledge, J.; Engelhaupt, E.; Burow, M.E.; McLachlan, J.A. Pesticides reduce symbiotic efficiency of nitrogen-fixing rhizobia and host plants. Proc. Natl. Acad. Sci. USA 2007, 104, 10282–10287. [Google Scholar] [CrossRef] [PubMed]
  109. Spokas, K.; Koskinen, W.; Baker, J.; Reicosky, D. 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]
  110. Castaldi, S.; Riondino, M.; Baronti, S.; Esposito, F.; Marzaioli, R.; Rutigliano, F.; Vaccari, F.; Miglietta, F. Impact of biochar application to a Mediterranean wheat crop on soil microbial activity and greenhouse gas fluxes. Chemosphere 2011, 85, 1464–1471. [Google Scholar] [CrossRef]
  111. Gibson, C.; Berry, T.D.; Wang, R.; Spencer, J.A.; Johnston, C.T.; Jiang, Y.; Bird, J.A.; Filley, T.R. Weathering of pyrogenic organic matter induces fungal oxidative enzyme response in single culture inoculation experiments. Org. Geochem. 2016, 92, 32–41. [Google Scholar] [CrossRef]
  112. Smith, K.S.; Balistrieri, L.S.; Smith, S.M.; Severson, R.C. Distribution and mobility of molybdenum in the terrestrial environment. In Molybdenum in Agriculture; Gupta, U.C., Ed.; Cambridge University Press: Cambridge, UK, 1997; pp. 23–46. [Google Scholar]
  113. Levy-Booth, D.J.; Prescott, C.E.; Grayston, S.J. Microbial functional genes involved in nitrogen fixation, nitrification and denitrification in forest ecosystems. Soil Biol. Biochem. 2014, 75, 11–25. [Google Scholar] [CrossRef]
  114. Nicol, G.W.; Leininger, S.; Schleper, C.; Prosser, J.I. The influence of soil pH on the diversity, abundance and transcriptional activity of ammonia oxidizing archaea and bacteria. Environ. Microbiol. 2008, 10, 2966–2978. [Google Scholar] [CrossRef]
  115. Yao, H.; Gao, Y.; Nicol, G.W.; Campbell, C.D.; Prosser, J.I.; Zhang, L.; Han, W.; Singh, B.K. Links between Ammonia Oxidizer Community Structure, Abundance, and Nitrification Potential in Acidic Soils. Appl. Environ. Microbiol. 2011, 77, 4618–4625. [Google Scholar] [CrossRef]
  116. Dempster, D.N.; Gleeson, D.; Solaiman, Z.; Jones, D.L.; Murphy, D. 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]
  117. 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. Soil Sci. Plant Nutr. 2007, 53, 181–188. [Google Scholar] [CrossRef]
  118. Borken, W.; Brumme, R. Liming practice in temperate forest ecosystems and the effects on CO2, N2O and CH4 fluxes. Soil Use Manag. 2007, 13, 251–257. [Google Scholar] [CrossRef]
  119. Enders, A.; Hanley, K.; Whitman, T.; Joseph, S.; Lehmann, J. Characterization of biochars to evaluate recalcitrance and agronomic performance. Bioresour. Technol. 2012, 114, 644–653. [Google Scholar] [CrossRef] [PubMed]
  120. Novak, J.M.; Busscher, W.J.; Laird, D.L.; Ahmedna, M.; Watts, D.W.; Niandou, M.A. Impact of Biochar Amendment on Fertility of a Southeastern Coastal Plain Soil N Volume. Soil Sci. 2009, 174, 105–112. [Google Scholar] [CrossRef]
  121. Tomczyk, A.; Sokołowska, Z.; Boguta, P. Biochar physicochemical properties: Pyrolysis temperature and feedstock kind effects. Rev. Environ. Sci. Bio/Technol. 2020, 19, 191–215. [Google Scholar] [CrossRef]
  122. Liang, B.; Lehmann, J.; Solomon, D.; Kinyangi, J.; Grossman, J.; O’Neill, B.J.; Skjemstad, J.O.; Thies, J.; Luizão, F.J.; Petersen, J.; et al. Black carbon increases cation exchange capacity in soils. Soil Sci. Soc. Am. J. 2006, 70, 1719. [Google Scholar] [CrossRef]
  123. 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]
  124. Antal, M.J.; Gronli, M. The art, science, and technology of charcoal production. Ind. Eng. Chem. Res. 2003, 42, 1619–1640. [Google Scholar] [CrossRef]
  125. Martinsen, V.; Alling, V.; Nurida, N.; Mulder, J.; Hale, S.; Ritz, C.; Rutherford, D.; Heikens, A.; Breedveld, G.; Cornelissen, G. pH effects of the addition of three biochars to acidic Indonesian mineral soils. Soil Sci. Plant Nutr. 2015, 61, 821–834. [Google Scholar] [CrossRef]
  126. Mia, S.; Dijkstra, F.A.; Singh, B. Enhanced biological nitrogen fixation and competitive advantage of legumes in mixed pastures diminish with biochar aging. Plant Soil 2018, 424, 639–651. [Google Scholar] [CrossRef]
  127. Wang, D.; Mukome, F.N.; Yan, D.; Wang, H.; Scow, K.M.; Parikh, S.J. Phenylurea herbicide sorption to biochars and agricultural soil. J. Environ. Sci. Health Part B 2015, 50, 544–551. [Google Scholar] [CrossRef] [PubMed]
  128. Lawrinenko, M.; Laird, D.A. Anion Exchange Capacity of Biochar. Green Chem. 2015, 17, 4628–4636. [Google Scholar] [CrossRef]
  129. Knowles, O.A.; Robinson, B.H.; Contangelo, A.; Clucas, L. Biochar for the mitigation of nitrate leaching from soil amended with biosolids. Sci. Total Environ. 2011, 409, 3206–3210. [Google Scholar] [CrossRef]
  130. Fabbri, D.; Rombolà, A.G.; Torri, C.; Spokas, K.A. Determination of polycyclic aromatic hydrocarbons in biochar and biochar amended soil. J. Anal. Appl. Pyrolysis 2013, 103, 60–67. [Google Scholar] [CrossRef]
  131. Lataf, A.; Jozefczak, M.; Vandecasteele, B.; Viaene, J.; Schreurs, S.; Carleer, R.; Yperman, J.; Marchal, W.; Cuypers, A.; Vandamme, D. The effect of pyrolysis temperature and feedstock on biochar agronomic properties. J. Anal. Appl. Pyrolysis 2022, 168, 105728. [Google Scholar] [CrossRef]
  132. Kim, E.J.; Oh, J.E.; Chang, Y.S. Effects of forest fire on the level and distribution of PCDD/Fs and PAHs in soil. Sci. Total Environ. 2003, 311, 177–189. [Google Scholar] [CrossRef]
  133. Wang, Z.; Zheng, H.; Luo, Y.; Deng, X.; Herbert, S.; Xing, B. Characterization and influence of biochars on nitrous oxide emission from agricultural soil. Environ. Pollut. 2013, 174, 289–296. [Google Scholar] [CrossRef]
  134. Wang, D.; Fonte, S.J.; Parikh, S.J.; Six, J.; Scow, K.M. Biochar additions can enhance soil structure and the physical stabilization of C in aggregates. Geoderma 2017, 303, 110–117. [Google Scholar] [CrossRef]
  135. Güereña, D.; Lehmann, J.; Hanley, K.; Enders, A.; Hyland, C.; Riha, S. Nitrogen dynamics following field application of biochar in a temperate North American maize-based production system. Plant Soil 2013, 365, 239–254. [Google Scholar] [CrossRef]
  136. Sarkhot, D.V.; Ghezzehei, T.A.; Berhe, A.A. Effectiveness of biochar for sorption of ammonium and phosphate from dairy effluent. J. Environ. Qual. 2013, 42, 1545–1554. [Google Scholar] [CrossRef] [PubMed]
  137. Domingues, R.R.; Sánchez-Monedero, M.A.; Spokas, K.A.; Melo, L.C.; Trugilho, P.F.; Valenciano, M.N.; Silva, C.A. Enhancing Cation Exchange Capacity of Weathered Soils Using Biochar: Feedstock, Pyrolysis Conditions and Addition Rate. Agronomy 2020, 10, 824. [Google Scholar] [CrossRef]
  138. Lin, Y.; Munroe, P.; Joseph, S.; Henderson, R.; Ziolkowski, A. Water extractable organic carbon in untreated and chemical treated biochars. Chemosphere 2012, 87, 151–157. [Google Scholar] [CrossRef] [PubMed]
  139. 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]
  140. Ball, P.N.; MacKenzie, M.D.; DeLuca, T.H.; Montana, W.H. 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]
  141. Harter, J.; Krause, H.M.; Schuettler, S.; Ruser, R.; Fromme, M.; Scholten, T.; Kappler, A.; Behrens, S. Linking N2O emissions from biochar-amended soil to the structure and function of the N-cycling microbial community. Int. Soc. Microb. Ecol. 2013, 8, 660–674. [Google Scholar] [CrossRef]
Figure 1. Simplified nitrogen cycle.
Figure 1. Simplified nitrogen cycle.
Sustainability 15 16446 g001
Figure 2. Potential mechanisms of biochar influence on fixation of atmospheric N2 and production of inorganic nitrogen (arrows denote change in product formation rate).
Figure 2. Potential mechanisms of biochar influence on fixation of atmospheric N2 and production of inorganic nitrogen (arrows denote change in product formation rate).
Sustainability 15 16446 g002
Figure 3. Potential mechanisms of the influence of biochar on assimilation of inorganic nitrogen and production of organic nitrogen, e.g., amino acids (arrows denote change in product formation rate).
Figure 3. Potential mechanisms of the influence of biochar on assimilation of inorganic nitrogen and production of organic nitrogen, e.g., amino acids (arrows denote change in product formation rate).
Sustainability 15 16446 g003
Figure 4. Potential mechanisms of the influence of biochar on mineralisation of organic nitrogen to non-organic forms (arrows denote change in product formation rate).
Figure 4. Potential mechanisms of the influence of biochar on mineralisation of organic nitrogen to non-organic forms (arrows denote change in product formation rate).
Sustainability 15 16446 g004
Figure 5. Potential mechanisms of the influence of biochar on denitrification of inorganic nitrogen (arrows denote change in product formation rate).
Figure 5. Potential mechanisms of the influence of biochar on denitrification of inorganic nitrogen (arrows denote change in product formation rate).
Sustainability 15 16446 g005
Table 1. Selected physical and chemical parameters for a range of biochars.
Table 1. Selected physical and chemical parameters for a range of biochars.
Biochar FeedstockC
g kg−1
N
g kg−1
C:N RatioP
g kg−1
K
g kg−1
pHPorosity Surface Area M2 g−1Reference
Wood chips70810.9656.80.9 4.82[16]
Wood chips72010.8671.3 9 [17]
Green wastes6801.74000.21.0 2.10[18]
Poultry litter380191925.222.1 [18]
Cow manure 20 2802010 [19]
Maize cobs5680.8710 18.711 [20]
Maize straw48912.539 104[21]
Pig manure51121.12438.5 10 [17]
Swine manure4224 30810 [22]
Birds foot trefoil foliage60032.218 9 [23]
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

Davys, D.; Rayns, F.; Charlesworth, S.; Lillywhite, R. The Effect of Different Biochar Characteristics on Soil Nitrogen Transformation Processes: A Review. Sustainability 2023, 15, 16446. https://doi.org/10.3390/su152316446

AMA Style

Davys D, Rayns F, Charlesworth S, Lillywhite R. The Effect of Different Biochar Characteristics on Soil Nitrogen Transformation Processes: A Review. Sustainability. 2023; 15(23):16446. https://doi.org/10.3390/su152316446

Chicago/Turabian Style

Davys, Donna, Francis Rayns, Susanne Charlesworth, and Robert Lillywhite. 2023. "The Effect of Different Biochar Characteristics on Soil Nitrogen Transformation Processes: A Review" Sustainability 15, no. 23: 16446. https://doi.org/10.3390/su152316446

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