Next Article in Journal
Anaerobic Digestate from Biogas Plants—Nuisance Waste or Valuable Product?
Next Article in Special Issue
Preliminary Investigation of Possible Biochar Use as Carbon Source in Polyacrylonitrile Electrospun Fiber Production
Previous Article in Journal
The San Giovanni Baptistery in Florence (Italy): Assessment of the State of Conservation of Surfaces and Characterization of Stone Materials
Previous Article in Special Issue
Effect of Pyrolysis Temperature on Copper Aqueous Removal Capability of Biochar Derived from the Kelp Macrocystis pyrifera
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Characteristics and Applications of Biochar in Soil–Plant Systems: A Short Review of Benefits and Potential Drawbacks

1
Institute of Food Science and Technology, Hungarian University of Agriculture and Life Sciences, 1118 Budapest, Hungary
2
Geographical Institute, Research Centre for Astronomy and Earth Sciences, 1112 Budapest, Hungary
3
Department of Agri-Environmental Studies, Szent Istvan University, 1118 Budapest, Hungary
*
Author to whom correspondence should be addressed.
Appl. Sci. 2022, 12(8), 4051; https://doi.org/10.3390/app12084051
Submission received: 4 March 2022 / Revised: 12 April 2022 / Accepted: 14 April 2022 / Published: 16 April 2022
(This article belongs to the Special Issue Biochar: Preparation and Surface Adsorption Applications)

Abstract

:
The available literary data suggest the general applicability and benefits of different biochar products in various soil–plant–environment systems. Due to its high porosity, biochar might generally improve the physicochemical and biological properties of supplemented soils. Among the direct and indirect effects are (i) improved water-retention capacity, (ii) enhanced soil organic matter content, (iii) pH increase, (iv) better N and P availability, and (v) greater potential uptake of meso- and micronutrients. These are connected to the advantage of an enhanced soil oxygen content. The large porous surface area of biochar might indirectly protect the survival of microorganisms, while the adsorbed organic materials may improve the growth of both bacteria and fungi. On the other hand, N2-fixing Rhizobium bacteria and P-mobilizing mycorrhiza fungi might respond negatively to biochar’s application. In arid circumstances with limited water and nutrient availability, a synergistic positive effect was found in biochar–microbial combined applications. Biochar seems to be a valuable soil supplement if its application is connected with optimized soil–plant–environment conditions. This work aims to give a general review of the potential benefits and drawbacks of biochar application to soil, highlighting its impacts on the soil–plant–microbe system.

1. Introduction

The nearly one meter deep black soil of the Amazon River in South America is rich in carbon, calcium, magnesium, phosphorus, and trace elements. It is estimated that 10% of this area is covered by the so-called “black soil” or “Terra Preta” soil [1]. The main reason for its productivity is the dead biomass that has undergone thermal decomposition in an oxygen-free condition. The tropical climate can positively influence soil properties, transforming them in a favorable direction over time, without deleterious structural changes. Columbian soils with variable contents of biochar and different layers of carbon content are compared to control soils in Table 1. The fertile layers in those soils are much deeper (between 43 and 210 cm) than the 20 cm fertile layer in control soils nearby. The pH values in general are neutral or less acidic in comparison with the more acidic control soils without biochar amendment. Clay contents in these soils also tend to be greater, which could be related to the enhanced biological activity. The main reasons for the detected positive effects are the better aeration and improved water content of soils, as well as the indirect effects of greater microbiological abundance and activity, which might improve plant nutrition [2,3].
According to literary data, the Terra Preta soil is suitable for extracting carbon from the carbon cycle over a long period (centuries) and, in this way, the carbon-sequestration potential of soil can mitigate global warming [4]. The same effect was also realized in soils in which studied ancient biochar production. The carbon in these soils was sequestered and nutrients were absorbed for more than 30 years [5].
Terra Preta has served as a model for soil amendment and biochar industrial products are now used worldwide. During the preparation of biochar, the carbon content of mainly biomass-based feedstock materials is converted to aromatic carbon compounds and amorphous and graphitic-type structures during the process of so-called pyrolysis. By the end of the process, about 50% of the carbon content of the original biomass remains in the end-product. Still, this ratio is highly dependent on the conditions of the pyrolysis process [6]. The product created during such a carbonization process is referred to in the literature as “biochar,” referring to its biological, natural bio-origin [7]. Varying the temperature and duration of pyrolysis can create significant differences in the properties of the biochar, even when starting from the same ingredient materials [8]. Three different fractions can be formed during the process. One is the persistent fraction, which remains largely resistant to further biological or chemical degradation for quite a long period [5,9]. This fraction typically has an amorphous structure that can contain graphite-like crystalline structures [10] consisting of conjugated aromatic structures to which many different functional groups are attached [11]. Toxic molecules (PAHs) also arise in this group of compounds, which must be considered at the point of application [12]. The labile fraction is the second characteristic group of substances. In the case of plant-derived biochar, it consists mainly of cellulose and hemicellulose which have not been converted during pyrolysis. When the pyrolysis process is fast, the carbonization is imperfect, which might result in an increase in the proportion of this group of substances. Once biochar is applied to the soil, this fraction is mineralized relatively quickly, in as little as a few weeks, and its nutrients are made available for crops [13,14,15]. The third fraction of biochar is the ash fraction, which is the oxidized residue of the mineral element in biomass. Its ratio increases with the pyrolysis temperature and the amount of oxygen involved in the process. A higher ash content increases the pH of the biochar and the concentrations of readily soluble elements present in a mineralized form [16]. Recognizing its beneficial effects, biochar production from various organic “wastes” and byproducts of agricultural and industrial origin has become a promising research area, and there are now a wide variety of related products and applications. Biochar is the umbrella term for products of biological origin, but we can also distinguish “pyrochar” and “hydrochar” products, depending on the production methodology and raw materials used, with industrial waste materials occasionally involved as well. Most examples of pyrochar and hydrochar, although also called “biochar”, are not of biological origin (Figure 1).
Due to its phytonutrient composition, soil is suggested as the main application for biologically based biochar products, which are known to have significant amounts of humic and fulvic acids attached to their porous structure surfaces [17].
The question arises as to whether the effects of biochar are beneficial in all cases for the soil–plant–microbial relationship; are there any time and quantity limits to biochar application? Furthermore, what are the most influential soil-environmental factors for its application? Do the effects appear directly through the physicochemical properties of the soils or indirectly in the soil–plant microbial system? The purpose of this summary is to review the general aspects (benefits and some drawbacks) of biochar application that might support its proper use in variable soil-environmental conditions, particularly to support the better growth of plants.

2. Approach to the Concept and Production of Biochar

Biochar is a substance formed via the reductive pyrolysis of high-carbon-containing dead plant and/or animal biomass (heated oxygen-free at high temperatures). Its structural properties strongly depend on the organic materials used for the reductive “burn”, formation temperature, and duration of action, i.e., the production conditions. The literature distinguishes two major groups of production processes in terms of the initial raw materials:
  • Biochar produced a relatively low (450–550 °C) pyrolyzing temperature, most often produced with high carbon-content substances according to the literature. It is mainly capable of long-term binding of groundwater and dissolved ions, and usually originating from plant residues, byproducts, and/or animal manures.
  • Biochar produced from animal bones at high temperatures (600–650 °C or higher) with a high calcium phosphate content, along with apatite minerals with significantly lower carbon contents.
In addition to biochar, oils and various synthesis gases are also generated during pyrolysis (Table 2). However, their exact proportions in the final products may vary as a function of the raw materials and production methods.
“Terra Preta” soil is well supplied with both organic matter and minerals, the most important of which are nitrogen, phosphorus, and potassium (NPP macronutrients) (Table 1). There are both direct and indirect reasons for this. In biochar production, plant waste and animal manure are mainly used, and the final product’s composition is proportional to the initial concentrations of raw materials. Its morphology is closely related to pore development, as well as to the shapes of pores. The smaller the pores, the larger the biochar surface area in general [22]. Biochar can show some variability in morphological structures when different initial feedstocks are used (Figure 2).
The biochar particle size distribution, structure, and mechanical resistance result in different effects under varying soil, climatic, and application conditions. Therefore, preliminary experiments should be conducted to optimize biochar parameters. The quality of biochar can be generalized within certain limits, considering the capacity of a given facility and the raw materials. The weight of the end-product can be maximized when the quantities of emitted gases, vapors, industrial water, and other pollutants are kept to a minimum [27,28]. Technological optimization is a long process, which is made more difficult by the almost constant composition of the raw materials and the frequent need for maintenance due to the resulting oils. Due to these difficulties, “two-channel” production devices are now used, where production can be continuously upheld in one unit during maintenance of the other side. However, putting this into practice is not yet particularly cost-effective [28].

3. Physicochemical Properties of Soils Affected by Biochar

When mixed with soil, biochar can change the soil’s texture, pore size distribution, and bulk density, positively affecting the aeration and the water-holding capacity [29,30,31]. The final soil characteristics following biochar addition are highly dependent on the initial quality of the soil, and its type and composition (porosity) in general. The application of biochar can serve first to improve the soil physical status, which may through several indirect effects affect other soil chemical and biological characteristics. In clayey soils, for instance, a larger pore size can be suggested (larger than the main pores of the clay, as 10 nm in diameter) in order to increase the aeration of these rather compacted soils. The opposite is possible in soils with high sand contents, where biochar can be used to improve the water-holding capacity and nutrient adsorption, not aeration. The variability of the porous structure of biochar, therefore, can increase the amount of water stored and the size and distribution of minerals as soil formers [32]. Biochar can have a surface area of up to 800–5000 mm2 per gram, and this porous, airy structure indirectly contributes to a generally high microbial activity and surviving ability. This beneficial property of biochar is also exploited in contaminated soils where environmental (stress) conditions (air limitations, lack of nutrients, critically low temperatures) would otherwise impede the remediation processes, such as the degradation of antifreeze propylene glycol [33]. Meanwhile, in the case of other forms of pollution, the positive effect of biochar on soil porosity is highlighted. Still, microorganism habitats and the protective surface (for biofilm production) provided for microorganisms are also a key issue in many soils, especially colloid-poor sandy soils [33,34].
Biochar has a liming effect [35] and increases soil cation exchange capacity by raising the pH, influencing the availability of nutrients, and thus preventing their leaching [36,37]. For example, the availability of phosphate, iron, boron, zinc, and manganese has been shown to decrease at high pH [37]. Increasing pH also increases microbial nitrification, which results in losses of nitrate and limited availability of ammonium, the preferred nitrogen source for plants [38]. In this way, in some cases, the pH-raising effect of biochar can create unfavorable conditions for plants, particularly in calcareous soils; this can result in yield losses [38,39]. Biochar has its own nutrient content (Table 3); depending on the raw material used and the pyrolysis temperature, it may contain large amounts of macro- and micronutrients with different levels of availability [40]. Therefore, biochar can be directly involved in the nutrient replenishment of plants [41]; however, aging, oxidation, and microbial degradation can alter the surface functional groups and chemical structures of biochar, with the result that the sorption capacity of biochar generally decreases with time. The effect of the aging, oxidation, or microbial degradation of biochar-derived dissolved organic matter on pollutants and the soil environment over time is still unclear [42]. In biochar, nutrients can be present in three forms, depending on the production conditions and the rate of utilization, as mentioned above. The nutrients in the ash fraction can be taken up over short periods by plants, particularly in well-watered soils, resulting in known benefits for plant growth. Nutrients in the labile fraction become available only after a few weeks of mineralization, as a consequence of the enhanced soil’s biological activity. Nutrients incorporated into the bodies of microorganisms become unavailable for a certain period, which may produce competition for nutrients between plants and microorganisms. This fact can result a so-called pentozan effect, which can develop in soils with very low nitrogen content but which are high in degradable organic materials (e.g., wheat straw) [43]. The nutrients in the persistent fraction of biochar are only released decades or centuries later, as discussed earlier [9]. These fractions can all play a role in the direct nutrient supply of plants.
An example of the direct, rapid nutrient replenishment of biochar was described by Angst et al. in 2013 [41]. In their study, the magnitude and dynamics of short-term phosphorus (P), magnesium (Mg), and potassium (K) release were assessed through repeated cold-water extraction using hardwood biochar (Acer pseudoplatanus). The feedstock was converted in a traditional charcoal ring furnace with a holding time of 24 h and a peak temperature of approximately 500 °C. The proximate analysis results for the char of volatiles, fixed carbon, and ash were 18.5%, 77.0%, and 4.4%, respectively. The carbon, hydrogen, oxygen, nitrogen, and sulfur contents were 746, 28, 178, 4.0, and 0.2 mg kg−1, respectively. The phosphorus concentration in the biochar (119 mg kg−1) was approximately four times greater than that in the Acer spp. feedstock, reflecting the ratio of feedstock mass to biochar yield in biochar production. In contrast, the Mg and K contents of 1889 and 3309 mg kg−1 were twice those of the feedstock. Approximately 30–103% of the total P, 6–27% of the total Mg, and 82–122% of the total K were leached from the biochar.
Increased nutrient retention is a vital element of harmonious nutrient supply. Biochar soil treatment significantly reduces the leaching of nutrients from organic and inorganic fertilizers [52]. In their nutrient leaching experiment in 2020, Laird et al. [53] significantly reduced the N, P, Mg, and Si contents of water-leached clay soils treated with organic manure. In summary, the beneficial effects of biochar on the soil’s ability to provide nutrients to plants stem from the following factors: (1) readily soluble nutrients brought into the soil [15] and nutrients from the mineralization of the labile fraction [13,14,15]; (2) reduced leaching due to biochar’s high cation exchange capacity [54,55]; (3) the reduction of ammonia volatilization and lower N loss due to the removal of N2 and N2O emissions from denitrification [56,57]; and (4) N, P, and S retention due to increased biological activity [58,59]. It is important to note that some of the above processes are inseparable from soil biological functions.
Adding biochar creates a darker soil surface, which increases the soil temperature, thus increasing the microbial biomass and activity values, which is also a positive factor [58]. The effects on the improvement of light absorption capacity and the increase of soil temperature can be efficiently utilized in the agricultural practices of boreal countries. Biochar usage can thus extend, for example, the northern border of wheat and maize cultivation [59].

4. Biological Properties of Soils Affected by Biochar

Several studies have reported beneficial effects of biochar substances—artificially created by charring biological waste and byproducts—on soil productivity in terms of increasing biological activity and the sequestration of greenhouse gases (GHGs) [60,61,62]. Among GHGs, the reduction of nitrous oxide (N2O) and methane (CH4) has been highlighted [54,57,62], but carbon sequestration (C) is also a critical property [52]. Biochar as a substance is mainly recommended for acidic soils as it alkalizes the pH and improves nutrient retention through cation adsorption, thereby benefiting soil productivity (Table 3). However, the variety of experiments and sites described in the literature make it difficult to summarize the effects on soil biomass and draw appropriate conclusions. In many cases, the results are contradictory and highlight only one factor of a multifactor system. The conditions for applying biochar depend on the properties of the soil, environmental conditions, raw material, dose, and many other biotic and abiotic factors. Its usage can lead to changes in soil biotic communities that are of interest but also of concern. In the quantitative development of the microbial community, not only plant-growth-promoting rhizobacteria (PGPR), which are favorable for cultivation, but also soil-borne pathogens may potentially multiply [63,64]. From the point of view of soil protection, such research is important, as the microbiological communities of soil affect both its functions and ecosystem services [65,66,67]. The properties that affect functionality are structure and stability, aeration, nutrient cycling, water balance, and the already mentioned carbon sequestration [68]. This refers to soils’ suppressive or receptive properties against pathogens, to which biochar can contribute in both positive and negative ways.
Overall, increased microbial biomass following biochar supplementation has been determined with various soil-testing methods. The most widely known of these are molecular nucleic-acid-based techniques [69,70], breeding, and/or the classical colony-counting method [64,71], substrate-induced respiration [72,73,74], fumigation extraction [74], phospholipid fatty acid (PLFA) analysis [70], and microscopic examination of staining particles [75,76,77]. In addition, Tian et al. in 2016 [78] and Wang et al. in 2020 [79] found that microbial reproduction rates also increased in biochar-treated soils. The increased microbial activity was further reflected in the CH4 formed during biodegradation after biochar addition. An increase was observed in the number of anaerobic and cellulose-degrading bacteria as an indirect effect. Thus, there may be a wide variety of microbes, of different types and functions in terms of plant production and protection, food quality, and safety, within each physiological group. The magnitude of these changes differs from species to species and even between strains [12]. Their increasing mass is a common phenomenon with biochar, but its extent and possible maximum are strongly determined by the ecophysiological properties of the studied taxon [80,81,82]. This has also been observed in stress-laden saline soils [83].
Biochar, like high-carbon sugars, immobilizes the easily absorbed nitrogen content of the soil and can thus be a successful tool in environmental restoration efforts [84]. In soil–plant systems, the microbiological properties of the rhizosphere have been mainly studied, as is shown in Table 4. Rékási et al. in 2010 [85] and Javeed et al. in 2019 [86] found that biochar delivered to the rhizosphere increases the number of microbes compared to that in undisturbed/untreated soil. However, generalization is prevented by the fact that Graber et al. [87] reported the opposite in 2010, referring to the strong dose-dependence of the application. Regarding microsymbionts, the presence and biomass of the two most common, arbuscular (AM) and ectomycorrhiza (EM), have also been shown to be positively influenced by the presence of biochar [75,76,83]. Both the rate of ectomycorrhizal formation and the level of root colonization increased (by 157%) at the roots of larch (Larix gmelinii) seedlings [80]. The AM colonization increased by 40% in a wheat rhizosphere two years after biochar application, while a 6% increase in mycorrhizal colonization was observed with 6 t ha−1 of biochar added to tree plantations [81]. However, we do not know how and to what extent the number of fungi and the extraradical hyphae mass extending beyond the root were affected by the added biochar. This is because the porous biochar can protect against physical damage resulting from soil compaction or a change to the nutrition of the soil fauna through its internal pores. Therefore, the actual surface protection of the microorganisms inside, consistent with their susceptibility to dehydration, is a critical issue in biochar treatment [12].
It has already been mentioned that one of the beneficial effects of biochar is on the surface binding of the gases produced in the soil, such as CO2. These indirect mechanisms influence the microbiological activity of soils and the composition of microbial groups [14,25,36,48,55]. However, the result is highly quality-dependent. In 2019, Rizwan [88] found that the properties of the hydrocarbon produced via the hydrothermal pathway differ significantly from the pyrolyzed version. On one hand, some authors have proposed that hydrothermal carbon better stimulates the germination and colonization of AM fungal spores in the soil due to its higher water content. On the other hand, other authors suggest that AM fungal colonization in hydrothermal char may decrease due to the improved physicochemical soil condition, which ultimately makes symbiosis unnecessary [88,89]. Decreased AM symbiosis has been observed with the use of high-phosphorus bone-derived biochar [90,91,92]. Thus, the essential role of beneficial mycorrhizal fungi that provide live, direct contact (their well-known properties that improve uptake of phosphorus and other elements, or even their stress-buffering capabilities) may be triggered by the use of various biochar products.
The direct adverse effects of biochar can even be demonstrated in soils with high salt or heavy metal contents, which inhibit the formation and function of the rhizosphere connection due to their poor chemical properties [93,94,95]. Depending on the surface and porosity of the biochar, it can bind not only the contaminants but also the organic nutrients essential for plant nutrition. The absence of these nutrients is of paramount importance in stress-laden soils [96,97]. In summary, biochar reduces the mutual efficiency of the plant–microbe system, which is an essential element of cooperation and symbiosis. According to previous results, in the absence of biochar, an improvement can be observed in the symbiotic relationships within soils contaminated with saline or heavy metals due to environmental stress [98,99]. In addition, biochar must be applied while considering the sensitivity of the symbiont relationship, which is why it should only be used with the appropriate expertise.
The beneficial effect of biochar can also be reduced by overwhelming the nutrient availability caused by the addition of fertilizer, which can indirectly reduce the growth rates of microbes [16,82,90]. Improvement in AM colonization was observed only with the smallest amount of “starter” fertilizer application [83]. According to Biró in 2000 [101], the nitrogen dose capable of promoting symbiosis in a low (1.5%) soil content corresponded to 45 kg ha−1 of nitrogen fertilizer and only 120 kg ha−1 of phosphorus. Even higher fertilizer doses caused multiple declines in symbiont functionality and economical biological N2 binding. The nutrient uptake effect of mycorrhizal fungi was reduced mainly by phosphorus-containing fertilizers but not by nitrogen-containing fertilizers [101]. In the initial development stage of pulse crops, when studying the nitrogen-binding Rhizobium bacteria capable of symbiotic contact, Ogawa and Okimori in 2010 [102] found the opposite result. With the addition of nitrogen and phosphorus fertilizer, symbiosis did not develop, or if it did, it was later downsized by the plant. In the existing symbiosis, the richness of the arbusculum that signals and ensures functioning can drop or re-emerge in as little as eight days, depending on the plant’s needs [83].
Under the same environmental conditions, the development of microbial biomass shows an increasing tendency to rise in the range from pH 3.7 to pH 8.3 [68,103,104]. However, fungi and bacteria react differently to changes in pH value. For example, the number of bacteria increases at around pH 7, while the fungal biomass does not change significantly at such neutral values [104]. A similar result was observed in Rhizobium-inoculated pulse crops [101,105]. Depending on the biochar raw material, its production temperature, and the amount applied, the soil’s acidity could be less than 4 or its alkalinity could be more than 12 [21,39,58,73]. The degree of biochar oxidation after entry into the soil plays an essential role in this process [7,11,14,72].
In summary, both the physicochemical adsorption properties of biochar (direct effect) and the mineralization due to microorganisms settling in more significant amounts in the pores of the biochar (indirect effect) contribute to an increase in plant nutrient uptake. Thus, the use of biochar does not inhibit but rather aids natural soil biological processes, which will allow the combined use of biochar and microbiological inoculants (biofertilizers, plant conditioners, other agents) in crop production in the future [38,50,52,60,69].

5. Soil Productivity Effects of Biochar

Most research results to date have confirmed biochar’s beneficial effect for increasing soil productivity, which is known as the most important soil function. The authors indicated a growth-inducing effect of biochar yield in almost 90% of the reviewed studies. The processed scientific publications reported yield increases of between 20% and 220% in proportion to the quantity and quality of biochar used. Some authors [106] considered only biochar application to intensive arable crops, and there were therefore no data available from settings such as organic farming areas. In an earlier study, Major [107] examined the increase and nutrient replenishment of a maize yield with soya sowing. Plant–coal biochar was applied at 8 and 20 t ha−1. The yield did not increase in the first year, but 28%, 30%, and 140% improvements were found in pots after applying a 20 t ha−1 dose in the following years. The effect was attributed to the better nutrient uptake, mainly via the soil’s 77–320% higher Ca and Mg contents in biochar-treated plots. At the same time, a more moderate increase in yield in the later years indicated a dose effect and the shortcomings of our current knowledge, which draws attention to the importance of long-term studies and the need for continuous field experiments and soil monitoring. Table 5 shows a summary of possible mechanisms by which physical, chemical, and biological properties can be affected by biochar supplementation in soil.

6. Conclusions

The major effects of biochar application on soil physical, chemical, and biological characteristics are summarized in Table 5. Regarding the objective of the present review, we have highlighted the benefits and some drawbacks of biochar application. Most of the literature shows that biochar generally contributes to improvement of physicochemical properties in soils, such as their water balance, clay, organic matter content, pH buffering, and the amounts of macro-, meso-, and several microelements, due to its porous structure, aeration, oxygen content, and relative quantity. The available literature indicates, therefore, a wide range of uses of biochar for the development and enhancement of beneficial soil–plant–microbial interactions. Improvements in plant nutrition may occur via several direct and indirect effects in soils. We have to mention, however, that intensification of soil biological activity is frequently a drawback of biochar application in some soils that are relatively low in soil organic matter and high in aeration, such as the arenosols. In arid environmental conditions, it is the limited water availability that is able to diminish known beneficial effects and reduce the potential nutrient uptake by crops. Integration of biochar into crop production technologies therefore requires preliminary experiments, particularly in considering the effects of symbiotic beneficial bacteria and fungi. The abundance and activity of microsymbionts such as the nitrogen-fixing Rhizobium bacteria and phosphorus-mobilizing mycorrhizal fungi can be dose- and product-dependent; thus, optimization seems to be necessary in order to suit specific soil–plant-environment circumstances. Further studies are suggested to efficiently address these drawbacks in biochar application.

Author Contributions

Conceptualization, project background, and advice, B.B.; data curation, formal analysis M.R.; funding acquisition, methodology, project administration, and writing original draft, T.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the European Union and co-financed by the European Social Fund (GA EFOP-3.6.3-VEKOP-16-2017-00005, VEKOP-2.3.3-15-2017-00022), EU, European Regional Development Fund, and Hungarian Government (GA GINOP-2.2.1-18-2020-00025); and EIP AGRI 1924228035 project.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author. The data are not publicly available due to privacy concerns.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ameloot, N.; Graber, E.R.; Verheijen, F.G.; De Neve, S. Interactions between biochar stability and soil organisms: Review and research needs. Eur. J. Soil Sci. 2013, 64, 379–390. [Google Scholar] [CrossRef]
  2. Ragályi, P.; Bernhardt, B.; Rékási, M.; Draskovits, E.; Molnár, S.; Molnár, M.; Uzinger, N. Effect of biochar and microbial inoculant on the element compostion and element yield of maize on acidic and on calcareous sandy soils. Agrokémia Talajt. 2019, 68, 115–137. [Google Scholar] [CrossRef]
  3. Pabar, S.A.; Mónok, D.; Kotroczó, Z.; Biró, B. Soil microbial parameters and synergies between bean growth and microbial inoculums as a dependence of five soils with different characteristics. Hung. Agric. Eng. 2020, 37, 27–33. [Google Scholar] [CrossRef]
  4. Lehmann, J.; Cowie, A.; Masiello, C.A.; Kammann, C.; Woolf, D.; Amonette, J.E.; Whitman, T. Biochar in climate change mitigation. Nat. Geosci. 2021, 14, 883–892. [Google Scholar] [CrossRef]
  5. Kocsis, T.; Biró, B.; Ulmer, Á.; Szántó, M.; Kotroczó, Z. Time-lapse effect of ancient plant coal biochar on some soil agrochemical parameters and soil characteristics. Environ. Sci. Pollut. Res. 2018, 25, 990–999. [Google Scholar] [CrossRef] [PubMed]
  6. Tripathi, N.; Hills, C.D.; Singh, R.S.; Atkinson, C.J. Biomass waste utilisation in low-carbon products: Harnessing a major potential resource. NPJ Clim. Atmos. Sci. 2019, 2, 1–10. [Google Scholar] [CrossRef] [Green Version]
  7. Luz, F.C.; Cordiner, S.; Manni, A.; Mulone, V.; Rocco, V. Biochar characteristics and early applications in anaerobic digestion-a review. J. Environ. Chem. Eng. 2018, 6, 2892–2909. [Google Scholar]
  8. Fuchs, M.R.; Garcia-Perez, M.; Small, P.; Flora, G. Campfire lessons: Breaking down the combustion process to understand biochar production and characterization. Biochar J. 2014. Available online: www.biochar-journal.org/en/ct/44 (accessed on 1 April 2022).
  9. Conte, P.; Nestle, N. Water dynamics in different biochar fractions. Magn. Reson. Chem. 2015, 53, 726–734. [Google Scholar] [CrossRef] [Green Version]
  10. Mohammad-Khah, A.; Ansari, R. Activated charcoal: Preparation, characterization and applications: A review article. Int. J. ChemTech Res. 2009, 1, 859–864. [Google Scholar]
  11. Kleber, M.; Hockaday, W.; Nico, P.S. Characteristics of biochar: Macro-molecular properties. In Biochar for Environmental Management; Routledge: London, UK, 2015; pp. 143–170. [Google Scholar]
  12. Kocsis, T.; Pabar, S.A.; Ferschl, B.; Kotroczó, Z.; Mohácsi-Farkas, C.; Biró, B. Biotic and abiotic risks of soil biochar treatment for food safety and human health. Acta Univ. Sapientiae Aliment. 2020, 13, 69–84. [Google Scholar] [CrossRef]
  13. Lawal, A.A.; Hassan, M.A.; Zakaria, M.R.; Yusoff, M.Z.M.; Norrrahim, M.N.F.; Mokhtar, M.N.; Shirai, Y. Effect of oil palm biomass cellulosic content on nanopore structure and adsorption capacity of biochar. Bioresour. Technol. 2021, 332, 125070. [Google Scholar] [CrossRef] [PubMed]
  14. Bakshi, S.; Banik, C.; Laird, D.A. Quantification and characterization of chemically-and thermally-labile and recalcitrant biochar fractions. Chemosphere 2018, 194, 247–255. [Google Scholar] [CrossRef] [PubMed]
  15. Yousaf, B.; Liu, G.; Abbas, Q.; Wang, R.; Ali, M.U.; Ullah, H.; Zhou, C. Systematic investigation on combustion characteristics and emission-reduction mechanism of potentially toxic elements in biomass-and biochar-coal co-combustion systems. Appl. Energy 2017, 208, 142–157. [Google Scholar] [CrossRef]
  16. Ye, L.; Camps-Arbestain, M.; Shen, Q.; Lehmann, J.; Singh, B.; Sabir, M. Biochar effects on crop yields with and without fertilizer: A meta-analysis of field studies using separate controls. Soil Use Manag. 2020, 36, 2–18. [Google Scholar] [CrossRef]
  17. Jindo, K.; Sonoki, T.; Matsumoto, K.; Canellas, L.; Roig, A.; Sanchez-Monedero, M.A. Influence of biochar addition on the humic substances of composting manures. Waste Manag. 2016, 49, 545–552. [Google Scholar] [CrossRef]
  18. Mao, J.D.; Johnson, R.L.; Lehmann, J.; Olk, D.C.; Neves, E.G.; Thompson, M.L.; Schmidt-Rohr, K. Abundant and stable char residues in soils: Implications for soil fertility and carbon sequestration. Environ. Sci. Technol. 2012, 46, 9571–9576. [Google Scholar] [CrossRef]
  19. 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]
  20. International Energy Agency. Annual Report—IEA Bioenergy. Task 34 Pyrolysis of Biomass. 2006. Available online: http://www.globalbioenergy.org/uploads/media/0707_IEA_-_Bioenergy_annual_report.pdf (accessed on 1 April 2022).
  21. Rehrah, D.; Reddy, M.R.; Novak, J.M.; Bansode, R.R.; Schimmel, K.A.; Yu, J.; Ahmedna, M. Production and characterization of biochars from agricultural by-products for use in soil quality enhancement. J. Anal. Appl. Pyrolysis 2014, 108, 301–309. [Google Scholar] [CrossRef]
  22. Leng, L.; Xiong, Q.; Yang, L.; Li, H.; Zhou, Y.; Zhang, W.; Huang, H. An overview on engineering the surface area and porosity of biochar. Sci. Total Environ. 2021, 763, 144204. [Google Scholar] [CrossRef]
  23. Abdel-Fattah, T.M.; Mahmoud, M.E.; Ahmed, S.B.; Huff, M.D.; Lee, J.W.; Kumar, S. Biochar from woody biomass for removing metal contaminants and carbon sequestration. J. Ind. Eng. Chem. 2015, 22, 103–109. [Google Scholar] [CrossRef]
  24. Côrtes, L.N.; Druzian, S.P.; Streit, A.F.M.; Godinho, M.; Perondi, D.; Collazzo, G.C.; Dotto, G.L. Biochars from animal wastes as alternative materials to treat colored effluents containing basic red 9. J. Environ. Chem. Eng. 2019, 7, 103446. [Google Scholar] [CrossRef]
  25. Pariyar, P.; Kumari, K.; Jain, M.K.; Jadhao, P.S. Evaluation of change in biochar properties derived from different feedstock and pyrolysis temperature for environmental and agricultural application. Sci. Total Environ. 2020, 713, 136433. [Google Scholar] [CrossRef] [PubMed]
  26. Demirbas, A. Production and characterization of bio-chars from biomass via pyrolysis. Energy Sources Part A 2006, 28, 413–422. [Google Scholar] [CrossRef]
  27. Gaunt, J.L.; Lehmann, J. Energy balance and emissions associated with biochar sequestration and pyrolysis bioenergy production. Environ. Sci. Technol. 2008, 42, 4152–4158. [Google Scholar] [CrossRef]
  28. Omara, P.; Aula, L.; Oyebiyi, F.B.; Eickhoff, E.M.; Carpenter, J.; Raun, W.R. Biochar application in combination with inorganic nitrogen improves maize grain yield, nitrogen uptake, and use efficiency in temperate soils. Agronomy 2020, 10, 1241. [Google Scholar] [CrossRef]
  29. Werner, S.; Akoto-Danso, E.K.; Manka’abusi, D.; Steiner, C.; Haering, V.; Nyarko, G.; Marschner, B. Nutrient balances with wastewater irrigation and biochar application in urban agriculture of Northern Ghana. Nutr. Cycl. Agroecosyst. 2019, 115, 249–262. [Google Scholar] [CrossRef]
  30. Verheijen, F.G.; Zhuravel, A.; Silva, F.C.; Amaro, A.; Ben-Hur, M.; Keizer, J.J. The influence of biochar particle size and concentration on bulk density and maximum water holding capacity of sandy vs sandy loam soil in a column experiment. Geoderma 2019, 347, 194–202. [Google Scholar] [CrossRef]
  31. Hussain, R.; Garg, A.; Ravi, K. Soil-biochar-plant interaction: Differences from the perspective of engineered and agricultural soils. Bull. Eng. Geol. Environ. 2020, 79, 4461–4481. [Google Scholar] [CrossRef]
  32. Rawal, A.; Joseph, S.D.; Hook, J.M.; Chia, C.H.; Munroe, P.R.; Donne, S.; Webber, J.B.W. Mineral–biochar composites: Molecular structure and porosity. Environ. Sci. Technol. 2016, 50, 7706–7714. [Google Scholar] [CrossRef] [Green Version]
  33. Libisch, B.; French, H.K.; Hartnik, T.; Anton, A.; Biró, B. Laboratory scale evaluation of selected remediation techniques for propylene-glycol based aircraft deicing fluids. Environ. Technol 2012, 33, 717–724. [Google Scholar] [CrossRef]
  34. Fekete, I.; Lajtha, K.; Kotroczó, Z.; Várbíró, G.; Varga, C.; Tóth, J.A.; Berki, I. Long-term effects of climate change on carbon storage and tree species composition in a dry deciduous forest. Glob. Chang. Biol. 2017, 23, 3154–3168. [Google Scholar] [CrossRef] [PubMed]
  35. Feigl, V.; Anton, A.; Uzigner, N.; Gruiz, K. Red mud as a chemical stabilizer for soil contaminated with toxic metals. Water Air Soil Pollut. 2012, 223, 1237–1247. [Google Scholar] [CrossRef]
  36. Fidel, R.B.; Laird, D.A.; Thompson, M.L.; Lawrinenko, M. Characterization and quantification of biochar alkalinity. Chemosphere 2017, 167, 367–373. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Cheng, H.; Jones, D.L.; Hill, P.; Bastami, M.S.; Tu, C.L. Influence of biochar produced from different pyrolysis temperature on nutrient retention and leaching. Arch. Agron. Soil Sci. 2018, 64, 850–859. [Google Scholar] [CrossRef]
  38. Xiao, L.; Meng, F. Evaluating the effect of biochar on salt leaching and nutrient retention of Yellow River Delta soil. Soil Use Manag. 2020, 36, 740–750. [Google Scholar] [CrossRef]
  39. Bachmann, H.J.; Bucheli, T.D.; Dieguez-Alonso, A.; Fabbri, D.; Knicker, H.; Schmidt, H.P.; Zehetner, F. Toward the standardization of biochar analysis: The COST action TD1107 interlaboratory comparison. J. Agric. Food Chem. 2016, 64, 513–527. [Google Scholar] [CrossRef]
  40. Wang, J.; Zhang, B.; Tian, Y.; Zhang, H.; Cheng, Y.; Zhang, J. A soil management strategy for ameliorating soil acidification and reducing nitrification in tea plantations. Eur. J. Soil Biol. 2018, 88, 36–40. [Google Scholar] [CrossRef]
  41. Joseph, S.; Cowie, A.L.; Van Zwieten, L.; Bolan, N.; Budai, A.; Buss, W.; Lehmann, J. How biochar works, and when it doesn’t: A review of mechanisms controlling soil and plant responses to biochar. GCB Bioenergy 2021, 13, 1731–1764. [Google Scholar] [CrossRef]
  42. Yang, F.; Wang, C.; Sun, H. A comprehensive review of biochar-derived dissolved matters in biochar application: Production, characteristics, and potential environmental effects and mechanisms. J. Environ. Chem. Eng. 2021, 9, 105258. [Google Scholar] [CrossRef]
  43. Kozyatnyk, I.; Oesterle, P.; Wurzer, C.; Mašek, O.; Jansson, S. Removal of contaminants of emerging concern from multicomponent systems using carbon dioxide activated biochar from lignocellulosic feedstocks. Bioresour. Technol. 2021, 340, 125561. [Google Scholar] [CrossRef] [PubMed]
  44. Jatav, H.S.; Rajput, V.D.; Minkina, T.; Singh, S.K.; Chejara, S.; Gorovtsov, A.; Kalinitchenko, V.P. Sustainable approach and safe use of biochar and its possible consequences. Sustainability 2021, 13, 10362. [Google Scholar] [CrossRef]
  45. Chan, K.Y.; Van Zwieten, L.; Meszaros, I.; Downie, A.; Joseph, S. Agronomic values of greenwaste biochar as a soil amendment. Aust. J. Soil Res. 2007, 45, 629–634. [Google Scholar] [CrossRef]
  46. Nguyen, B.; Lehmann, J.; Hockaday, W.C.; Joseph, S.; Masiello, C.A. Temperature sensitivity of black carbon decomposition and oxidation. Env. Sci. Technol. 2010, 44, 3324–3331. [Google Scholar] [CrossRef] [PubMed]
  47. Magrini-Bair, K.A.; Czernik, S.; Pilath, H.M.; Evans, R.J.; Maness, P.C.; Leventhal, J. Biomass derived, carbon sequestration, designed fertilizers. Ann. Environ. Sci. 2009, 3, 217–225. [Google Scholar]
  48. Prakongkep, N.; Gilkes, R.J.; Wiriyakitnateekul, W.; Duangchan, A.; Darunsontaya, T. The effects of pyrolysis conditions on the chemical and physical properties of rice husk biochar. Int. J. Mater. Sci. 2013, 3, 97–103. [Google Scholar]
  49. Forján, R.; Rodríguez-Vila, A.; Cerqueira, B.; Covelo, E.F. Comparison of the effects of compost versus compost and biochar on the recovery of a mine soil by improving the nutrient content. J. Geochem. Explor. 2017, 183, 46–57. [Google Scholar] [CrossRef]
  50. Yargicoglu, E.N.; Sadasivam, B.Y.; Reddy, K.R.; Spokas, K. Physical and chemical characterization of waste wood derived biochars. Waste Manag. 2015, 36, 256–268. [Google Scholar] [CrossRef]
  51. Zhang, H.; Voroney, R.P.; Price, G.W. Effects of temperature and processing conditions on biochar chemical properties and their influence on soil Cand N transformations. Soil Biol. Biochem. 2015, 83, 19–28. [Google Scholar] [CrossRef]
  52. Baigorri, R.; San Francisco, S.; Urrutia, Ó.; García-Mina, J.M. Biochar-Ca and biochar-Al/-Fe-mediated phosphate exchange capacity are main drivers of the different biochar effects on plants in acidic and alkaline soils. Agronomy 2020, 10, 968. [Google Scholar] [CrossRef]
  53. Laird, D.; Rogovska, N. Biochar effects on nutrient leaching. In Biochar for Environmental Management; Routledge: London, UK, 2015; pp. 553–574. [Google Scholar]
  54. Zhang, J.; Wang, Q. Sustainable mechanisms of biochar derived from brewers’ spent grain and sewage sludge for ammonia–nitrogen capture. J. Clean. Prod. 2016, 112, 3927–3934. [Google Scholar] [CrossRef]
  55. Borchard, N.; Schirrmann, M.; Cayuela, M.L.; Kammann, C.; Wrage-Mönnig, N.; Estavillo, J.M.; Novak, J. Biochar, Soil and land-use interactions that reduce nitrate leaching and N2O emissions: A meta-analysis. Sci. Total Environ. 2019, 651, 2354–2364. [Google Scholar] [CrossRef] [PubMed]
  56. Karimi, A.; Moezzi, A.; Chorom, M.; Enayatizamir, N. Application of biochar changed the status of nutrients and biological activity in a calcareous soil. J. Soil Sci. Plant Nutr. 2020, 20, 450–459. [Google Scholar] [CrossRef]
  57. Foster, E.J.; Baas, P.; Wallenstein, M.D.; Cotrufo, M.F. Precision biochar and inoculum applications shift bacterial community structure and increase specific nutrient availability and maize yield. Appl. Soil Ecol. 2020, 151, 103541. [Google Scholar] [CrossRef]
  58. Ding, Y.; Gao, X.; Qu, Z.; Jia, Y.; Hu, M.; Li, C. Effects of biochar application and irrigation methods on soil temperature in farmland. Water 2019, 11, 499. [Google Scholar] [CrossRef] [Green Version]
  59. Tammeorg, P.; Simojoki, A.; Mäkelä, P.; Stoddard, F.L.; Alakukku, L.; Helenius, J. Short-term effects of biochar on soil properties and wheat yield formation with meat bone meal and inorganic fertiliser on a boreal loamy sand. Agric. Ecosyst. Environ. 2014, 191, 108–116. [Google Scholar] [CrossRef]
  60. El-Naggar, A.; Lee, S.S.; Rinklebe, J.; Farooq, M.; Song, H.; Sarmah, A.K.; Ok, Y.S. Biochar application to low fertility soils: A review of current status, and future prospects. Geoderma 2019, 337, 536–554. [Google Scholar] [CrossRef]
  61. Fidel, R.B.; Laird, D.A.; Parkin, T.B. Effect of biochar on soil greenhouse gas emissions at the laboratory and field scales. Soil Syst. 2019, 3, 8. [Google Scholar] [CrossRef] [Green Version]
  62. Barracosa, P.; Cardoso, I.; Marques, F.; Pinto, A.; Oliveira, J.; Trindade, H.; Pereira, J.L. Effect of biochar on emission of greenhouse gases and productivity of cardoon crop (Cynara cardunculus L.). J. Soil Sci. Plant Nutr. 2020, 20, 1524–1531. [Google Scholar] [CrossRef]
  63. 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]
  64. Gurtler, J.B.; Boateng, A.A.; Han, Y.; Douds Jr, D.D. Inactivation of E. coli O157: H7 in cultivable soil by fast and slow pyrolysis-generated biochar. Foodborne Pathog. Dis. 2014, 11, 215–223. [Google Scholar] [PubMed]
  65. Kwon, T.; Shibata, H.; Kepfer-Rojas, S.; Schmidt, I.K.; Larsen, K.S.; Beier, C.; Gualmini, M. Effects of climate and atmospheric nitrogen deposition on early to mid-term stage litter decomposition across biomes. Front. For. Glob. Chang. 2021, 90, 1–18. [Google Scholar] [CrossRef]
  66. Nannipieri, P.; Ascher-Jenull, J.; Ceccherini, M.T.; Pietramellara, G.; Renella, G.; Schloter, M. Beyond microbial diversity for predicting soil functions: A mini review. Pedosphere 2020, 30, 5–17. [Google Scholar] [CrossRef]
  67. Nannipieri, P.; Ascher, J.; Ceccherini, M.T.; Landi, L.; Pietramellara, G.; Renella, G. Microbial diversity and soil functions. Eur. J. Soil Sci. 2017, 68, 12–26. [Google Scholar] [CrossRef]
  68. Tóth, J.A.; Nagy, P.T.; Krakomperger, Z.; Veres, Z.; Kotroczó, Z.; Kincses, S.; Fekete, I.; Papp, M.; Mészáros, I.; Viktor, O. The effects of climate change on element content and soil pH (Síkfőkút DIRT project, Northern Hungary). In The Carpathians: Integrating Nature and Society towards Sustainability. Environmental Science and Engineering; Kozak, J., Ostapowicz, K., Bytnerowicz, A., Wyżga, B., Eds.; Springer: Berlin/Heidelberg, Germany, 2013; pp. 77–88. [Google Scholar]
  69. Ennis, C.J.; Evans, A.G.; Islam, M.; Ralebitso-Senior, T.K.; Senior, E. Biochar: Carbon sequestration, land remediation, and impacts on soil microbiology. Crit. Rev. Environ. Sci. Technol. 2012, 42, 2311–2364. [Google Scholar] [CrossRef] [Green Version]
  70. Paz-Ferreiro, J.; Méndez, A.; Gascó, G.; Guo, M.; He, Z.; Uchimiya, S.M. Application of biochar for soil biological improvement. In Agricultural and Environmental Applications of Biochar: Advances and Barriers; Guo, M., He, Z., Uchimiya, S.M., Eds.; Wiley: Hoboken, NJ, USA, 2016; pp. 145–174. [Google Scholar]
  71. Zavalloni, C.; Alberti, G.; Biasiol, S.; Delle Vedove, G.; Fornasier, F.; Liu, J.; Peressotti, A. Microbial mineralization of biochar and wheat straw mixture in soil: A short-term study. Appl. Soil Ecol. 2011, 50, 45–51. [Google Scholar] [CrossRef]
  72. Smirnova, E.V.; Giniyatullin, K.G.; Okunev, R.; Valeeva, A.A.; Guseva, I.A. The effect of pre-incubation duration of soil-biochar model mixtures on the results of determination the intensity of substrate-induced respiration (methodological aspects of study). Res. J. Pharm. Biol. Chem. Sci. 2016, 7, 1360–1366. [Google Scholar]
  73. Bandara, T.; Franks, A.; Xu, J.; Chathurika, J.B.A.J.; Tang, C. Biochar aging alters the bioavailability of cadmium and microbial activity in acid contaminated soils. J. Hazard. Mater. 2021, 420, 126666. [Google Scholar] [CrossRef]
  74. Brtnicky, M.; Dokulilova, T.; Holatko, J.; Pecina, V.; Kintl, A.; Latal, O.; Datta, R. Long-term effects of biochar-based organic amendments on soil microbial parameters. Agronomy 2019, 9, 747. [Google Scholar] [CrossRef] [Green Version]
  75. Videgain-Marco, M.; Marco-Montori, P.; Martí-Dalmau, C.; Jaizme-Vega, M.D.C.; Manyà-Cervelló, J.J.; García-Ramos, F.J. The Effects of Biochar on Indigenous Arbuscular Mycorrhizae Fungi from Agroenvironments. Plants 2021, 10, 950. [Google Scholar] [CrossRef]
  76. Solaiman, Z.M.; Shafi, M.I.; Beamont, E.; Anawar, H.M. Poultry litter biochar increases mycorrhizal colonisation, soil fertility and cucumber yield in a fertigation system on sandy soil. Agriculture 2020, 10, 480. [Google Scholar] [CrossRef]
  77. Jaafar, N.M.; Clode, P.L.; Abbott, L.K. Soil microbial responses to biochars varying in particle size, surface and pore properties. Pedosphere 2015, 25, 770–780. [Google Scholar] [CrossRef]
  78. Tian, J.; Wang, J.; Dippold, M.; Gao, Y.; Blagodatskaya, E.; Kuzyakov, Y. Biochar affects soil organic matter cycling and microbial functions but does not alter microbial community structure in a paddy soil. Sci. Total Environ. 2016, 556, 89–97. [Google Scholar] [CrossRef] [PubMed]
  79. Wang, Y.; Zheng, J.; Liu, X.; Yan, Q.; Hu, Y. Short-term impact of fire-deposited charcoal on soil microbial community abundance and composition in a subtropical plantation in China. Geoderma 2020, 359, 113992. [Google Scholar] [CrossRef]
  80. 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]
  81. Solaiman, Z.M.; Blackwell, P.; Abbott, L.K.; Storer, P. Direct and residual effect of biochar application on mycorrhizal colonization, growth and nutrition of wheat. Austral. J. Soil Res. 2010, 48, 546–554. [Google Scholar] [CrossRef]
  82. Dangi, S.; Gao, S.; Duan, Y.; Wang, D. Soil microbial community structure affected by biochar and fertilizer sources. Appl. Soil Ecol. 2020, 150, 103452. [Google Scholar] [CrossRef]
  83. Füzy, A.; Biró, B.; Tóth, T. Effect of saline soil parameters on endomycorrhizal colonisation of dominant halophytes in four Hungarian sites. Span. J. Agric. Res. 2010, 1, 144–148. [Google Scholar] [CrossRef] [Green Version]
  84. Horel, Á.; Potyó, I.; Szili-Kovács, T.; Molnár, S. Potential nitrogen fixation changes under different land uses as influenced by seasons and biochar amendments. Arab. J. Geosci. 2018, 11, 1–10. [Google Scholar] [CrossRef] [Green Version]
  85. Rékási, M.; Szili-Kovács, T.; Takács, T.; Bernhardt, B.; Puspán, I.; Kovács, R.; Uzinger, N. Improving the fertility of sandy soils in the temperate region by combined biochar and microbial inoculant treatments. Arch. Agron. Soil Sci. 2019, 65, 44–57. [Google Scholar] [CrossRef]
  86. Javeed, H.M.R.; Qamar, R.; Ali, M.; Rehman, A.; Farooq, M.; Zamir, S.I.; Hussain, M. Improvement in Soil Characteristics of Sandy Loam Soil and Grain Quality of Spring Maize by Using Phosphorus Solublizing Bacteria. Sustainability 2019, 11, 7049. [Google Scholar] [CrossRef] [Green Version]
  87. Graber, E.R.; Harel, Y.M.; Kolton, M.; Cytryn, E.; Silber, A.; David, D.R.; Tsechansky, L.; Borenshtein, M.; Elad, Y. Biochar impact on development and productivity of pepper and tomato grown in fertigated soilless media. Plant Soil 2010, 337, 481–496. [Google Scholar] [CrossRef]
  88. Rizwan, M.; ur Rehman, M.Z.; Ali, S.; Abbas, T.; Maqbool, A.; Bashir, A. Biochar is a potential source of silicon fertilizer: An overview. Biochar Biomass Waste 2019, 12, 225–238. [Google Scholar]
  89. 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]
  90. Grafe, M.; Kurth, J.K.; Panten, K.; Raj, A.D.; Baum, C.; Zimmer, D.; Schulz, S. Effects of different innovative bone char-based P fertilizers on bacteria catalyzing P turnover in agricultural soils. Agric. Ecosyst. Environ. 2021, 314, 107419. [Google Scholar] [CrossRef]
  91. Nigri, E.M.; Santos, A.L.; Mesquita, P.L.; Viana, P.R.; Rocha, S.D. Simultaneous removal of strontium and refractory organic compounds from electrodialysis effluents by modified bovine bone char. Desalin. Water Treat. 2019, 145, 189–201. [Google Scholar] [CrossRef]
  92. Zwetsloot, M.J.; Lehmann, J.; Bauerle, T.; Vanek, S.; Hestrin, R.; Nigussie, A. Phosphorus availability from bone char in a P-fixing soil influenced by root-mycorrhizae-biochar interactions. Plant Soil 2016, 408, 95–105. [Google Scholar] [CrossRef]
  93. Tang, J.; Zhang, S.; Zhang, X.; Chen, J.; He, X.; Zhang, Q. Effects of pyrolysis temperature on soil-plant-microbe responses to Solidago canadensis L.-derived biochar in coastal saline-alkali soil. Sci. Total Environ. 2020, 731, 138938. [Google Scholar] [CrossRef]
  94. Inyang, M.I.; Gao, B.; Yao, Y.; Xue, Y.; Zimmerman, A.; Mosa, A.; Cao, X. A review of biochar as a low-cost adsorbent for aqueous heavy metal removal. Crit. Rev. Environ. Sci. Technol. 2016, 46, 406–433. [Google Scholar] [CrossRef]
  95. Komkiene, J.; Baltrenaite, E. Biochar as adsorbent for removal of heavy metal ions [Cadmium (II), Copper (II), Lead (II), Zinc (II)] from aqueous phase. Int. J. Environ. Sci. Technol. 2016, 13, 471–482. [Google Scholar] [CrossRef] [Green Version]
  96. El-Naggar, A.H.; Usman, A.R.; Al-Omran, A.; Ok, Y.S.; Ahmad, M.; Al-Wabel, M.I. Carbon mineralization and nutrient availability in calcareous sandy soils amended with woody waste biochar. Chemosphere 2015, 138, 67–73. [Google Scholar] [CrossRef] [PubMed]
  97. 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] [Green Version]
  98. Varma, A.; Tripathi, S.; Prasad, R. (Eds.) Plant Microbe Symbiosis; Springer Nature: Berlin/Heidelberg, Germany, 2020. [Google Scholar]
  99. Khatabi, B.; Gharechahi, J.; Ghaffari, M.R.; Liu, D.; Haynes, P.A.; McKay, M.J.; Salekdeh, G.H. Plant–microbe symbiosis: What has proteomics taught us? Proteomics 2019, 19, 1800105. [Google Scholar] [CrossRef] [PubMed]
  100. Lehmann, J.; Rillig, M.C.; Thies, J.; Masiello, C.A.; Hockaday, W.C.; Crowley, D. Biochar effects on soil biota. Soil Biol. Biochem. 2011, 43, 1812–1836. [Google Scholar] [CrossRef]
  101. Biró, B.; Köves-Péchy, K.; Vörös, I.; Takács, T.; Eggenberger, P.; Strasser, R.J. Interrelations between Azospirillum and Rhizobium nitrogen-fixers and arbuscular mycorrhizal fungi in the rhizosphere of alfalfa in sterile, AM-free or normal soil conditions. Appl. Soil Ecol. 2000, 15, 159–168. [Google Scholar] [CrossRef]
  102. Ogawa, M.; Okimori, Y. Pioneering works in biochar research, Japan. Soil Res. 2010, 48, 489–500. [Google Scholar] [CrossRef] [Green Version]
  103. Yu, L.; Yu, M.; Lu, X.; Tang, C.; Liu, X.; Brookes, P.C.; Xu, J. Combined application of biochar and nitrogen fertilizer benefits nitrogen retention in the rhizosphere of soybean by increasing microbial biomass but not altering microbial community structure. Sci. Total Environ. 2018, 640, 1221–1230. [Google Scholar] [CrossRef]
  104. Zhalnina, K.; Dias, R.; de Quadros, P.D.; Davis-Richardson, A.; Camargo, F.A.; Clark, I.M.; Triplett, E.W. Soil pH determines microbial diversity and composition in the park grass experiment. Microb. Ecol. 2015, 69, 395–406. [Google Scholar] [CrossRef]
  105. Shamseldin, A.; Werner, D. High salt and high pH tolerance of new isolated Rhizobium etli strains from Egyptian soils. Curr. Microbiol. 2005, 50, 11–16. [Google Scholar] [CrossRef]
  106. McHenry, M.P. Soil organic carbon, biochar, and applicable research results for increasing farm productivity under Australian agricultural conditions. Commun. Soil Sci. Plant Anal. 2011, 42, 1187–1199. [Google Scholar] [CrossRef]
  107. Major, J.; Rondon, M.; Molina, D.; Riha, S.J.; Lehmann, J. Maize yield and nutrition during 4 years after biochar application to a Colombian savanna oxisol. Plant Soil 2010, 333, 117–128. [Google Scholar] [CrossRef]
Figure 1. Main types of biochar and their agricultural uses.
Figure 1. Main types of biochar and their agricultural uses.
Applsci 12 04051 g001
Figure 2. Surface morphology of some biochar, based on the different feedstocks [22,23,24,25,26]. (A) pinewood-derived, (B) sawdust-derived, (C) bovine-bone-derived, (D) fish-scale-derived, (E) food-waste-derived, (F) poultry-litter-derived biochar.
Figure 2. Surface morphology of some biochar, based on the different feedstocks [22,23,24,25,26]. (A) pinewood-derived, (B) sawdust-derived, (C) bovine-bone-derived, (D) fish-scale-derived, (E) food-waste-derived, (F) poultry-litter-derived biochar.
Applsci 12 04051 g002
Table 1. Main physicochemical properties of the Columbian “Terra Preta” and similar non-biochar-affected control soils [18,19].
Table 1. Main physicochemical properties of the Columbian “Terra Preta” and similar non-biochar-affected control soils [18,19].
SiteSoilFertile Layer (cm)Age (year)Clay (%)pHCarbon (mg g−1)Nitrogen (mg g−1)C:N
Ratio
HataharaTerra Preta43–69600–100027.06.422.01.022
Control0–10600–100035.94.621.81.614
Lago GrandeTerra Preta0–16900–110022.65.931.51.818
Control0–8900–110026.74.217.51.313
AcutubaTerra Preta48–832000–300010.45.615.71.016
Control0–302000–30008.54.715.40.819
Dona StellaTerra Preta190–2106700–87000.35.016.51.115
Control0–126700–87000.33.910.20.426
Table 2. The products of initial feedstock mass under selected pyrolysis conditions [20,21].
Table 2. The products of initial feedstock mass under selected pyrolysis conditions [20,21].
Method of PyrolysisLiquid
(Bio-Oil, Bio-Fuel)
Solid
(Biochar)
Gas
(Synthesis Gas, Bio-Gas)
Moderate temperature (~500 °C), short hot vapor residence time (<2 s)75%
(25% water)
12%13%
Moderate temperature (<500 °C), moderate hot vapor residence time (10–20 s)50%
(50% water)
20%30%
Slow temperature (~400 °C), very long solid residence time30%
(70% water)
35%35%
High temperature (>800 °C), long vapor residence time5% tar10%85%
Table 3. Chemical constituents of biochar produced from various feedstock sources under different production temperatures (based on Jatav et al., 2021 [44]).
Table 3. Chemical constituents of biochar produced from various feedstock sources under different production temperatures (based on Jatav et al., 2021 [44]).
Raw
Material
Production (°C)pHC
g kg−1
N
g kg−1
C:N
Ratio
P
g kg−1
K
g kg−1
Literature
Green waste4506.26801.74000.21Chan et al., 2007 [45]
Poultry litter4509.938020192522Chan et al., 2007 [45]
Residue
(Zea mays)
350 6759.373 10.4Nguyen and Lehmann, 2009 [46]
Residue
(Zea mays)
600 7909.286 6.7Nguyen and Lehmann, 2009 [46]
Peanut shell
(Arachis hypogaea)
400 49911450.66.2Margini-Bair et al., 2009 [47]
Rice husk370–5508.54705.9801.037.9Prakongkep et al., 2014 [48]
Wood
(Quercus ilex)
4009.96765.3128 3.2Forján et al., 2014 [49]
Pine wood
(Pinus spp.)
~4808.45323.7143 9.4Yargicoglu et al., 2015 [50]
Wood
(Quercus spp.)
4006.94273.31300.63.8Zhang et al., 2015 [51]
Wood
(Quercus spp.)
6009.54554.11110.64.4Zhang et al., 2015 [51]
Table 4. Summary of possible mechanisms by which microbial abundance is affected by biochar supplementation of soil. (Original data from Lehmann et al., 2011 [100], were refreshed based on new results).
Table 4. Summary of possible mechanisms by which microbial abundance is affected by biochar supplementation of soil. (Original data from Lehmann et al., 2011 [100], were refreshed based on new results).
MechanismRhizobiumBacteriaMycorrhizaFilamentous Fungi
Protecting surface0+++
Improved hydration++++
N availability+0+
P, Ca, Mg, K availability++
Micronutrient availability+++
pH increase++00
pH decrease00
Sorption of microorganisms+++
Biofilm formation++00
Sorption of dissolved organic matter as an energy source for microorganisms0+0+
+ positive effect, − negative effect, 0 no change.
Table 5. Summary of possible advantages and disadvantages of biochar application in various soil–plant systems with suggested solutions.
Table 5. Summary of possible advantages and disadvantages of biochar application in various soil–plant systems with suggested solutions.
Soil CharacteristicsAdvantages of BCDisadvantages of BCSuggestions to Treat
Soil physical conditions
Texture, porosity- BC can be used to improve soil quality [21,86]- Soil type is crucial in positive effect [29,30,31,32]
- Site-specific application needed [10,60]
- Previous selection and study needed to avoid improper use
Surface area, plant-nutrition- Adsorption and fixing of elements or leachable materials (e.g., nitrate) [37]
- Protects soil biota [33,34]
- Plant nutrition might be limited (e.g., in drought conditions) [38,39]- Consider the stressed environmental condition (watering, soil inoculation)
Aeration, better oxygenation- Supports aerobic processes, protects soil biota [2,3]- Potential of reduced SOM and humus content [42]- Use soil-dependent treatments, add organic materials
Soil chemistry
pH- Near-neutral conditions, better for the soil life [100]- Some nutritive elements become less available [36]- Consider current soil characteristics and act accordingly
SOM, humus, carbon- Sequesters carbon [4,5]
- Mitigates climate change [58,59]
- Indirect effect on soil biota might reduce SOM [63,64]- Proper use might be required
Toxic materials, heavy metals- Improved decontamination [16,81,89]
- Heavy metal adsorption and fixing [99,100]
- Potential accumulation of toxic compounds [5]- Inoculation by adapted microbes might improve remediation
Soil biology
Survival
of soil biota
- Large surfaces can provide niche [65,66,67]
- Drought protection and improved stress tolerance [70,84,88]
- Dependence on microbial physiological groups [98]
- Limitation in microbial distribution [12]
- Focus on soil-borne plant pathogens might be helpful
Activity
of soil biota
- Enhanced plant nutrition [94,96,97]- Competition with plants for nutrients [43]- Proper C:N ratio to avoid penthozan effect
- Optimization for specific soil–plant systems
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Kocsis, T.; Ringer, M.; Biró, B. Characteristics and Applications of Biochar in Soil–Plant Systems: A Short Review of Benefits and Potential Drawbacks. Appl. Sci. 2022, 12, 4051. https://doi.org/10.3390/app12084051

AMA Style

Kocsis T, Ringer M, Biró B. Characteristics and Applications of Biochar in Soil–Plant Systems: A Short Review of Benefits and Potential Drawbacks. Applied Sciences. 2022; 12(8):4051. https://doi.org/10.3390/app12084051

Chicago/Turabian Style

Kocsis, Tamás, Marianna Ringer, and Borbála Biró. 2022. "Characteristics and Applications of Biochar in Soil–Plant Systems: A Short Review of Benefits and Potential Drawbacks" Applied Sciences 12, no. 8: 4051. https://doi.org/10.3390/app12084051

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