Submitted Successfully!
To reward your contribution, here is a gift for you: A free trial for our video production service.
Thank you for your contribution! You can also upload a video entry or images related to this topic.
Version Summary Created by Modification Content Size Created at Operation
1 -- 3451 2023-08-23 11:53:24 |
2 Update references Meta information modification 3451 2023-08-24 03:14:25 |

Video Upload Options

Do you have a full video?

Confirm

Are you sure to Delete?
Cite
If you have any further questions, please contact Encyclopedia Editorial Office.
Tsolis, V.; Barouchas, P. The Effect of Biochar on Soil Properties. Encyclopedia. Available online: https://encyclopedia.pub/entry/48366 (accessed on 17 May 2024).
Tsolis V, Barouchas P. The Effect of Biochar on Soil Properties. Encyclopedia. Available at: https://encyclopedia.pub/entry/48366. Accessed May 17, 2024.
Tsolis, Vasileios, Pantelis Barouchas. "The Effect of Biochar on Soil Properties" Encyclopedia, https://encyclopedia.pub/entry/48366 (accessed May 17, 2024).
Tsolis, V., & Barouchas, P. (2023, August 23). The Effect of Biochar on Soil Properties. In Encyclopedia. https://encyclopedia.pub/entry/48366
Tsolis, Vasileios and Pantelis Barouchas. "The Effect of Biochar on Soil Properties." Encyclopedia. Web. 23 August, 2023.
The Effect of Biochar on Soil Properties
Edit

Biochar, a product of biomass pyrolysis, is recognized for its positive effects on soil fertility and carbon sequestration. Biochar acts as a soil conditioner, improving physical, chemical, and biological properties and enhancing soil fertility and crop yield. Furthermore, it aids in mitigating climate change by sequestering carbon dioxide. However, the long-term behavior of biochar and its interactions with various factors require further field research for optimal utilization, as the aging process of biochar in soil is complex, involving physical, chemical, and biological interactions that influence its impact on the agroecosystem.

biochar soil organic carbon diffuse reflectance spectroscopy

1. Introduction

Soil can be characterized as a natural, non-renewable, vital, multivariable, complex, regulatory resource, a significant driver of agriculture, a complex organic material [1], an environmental filter of nutrients [2], and a large pool of soil carbon [3] since its contribution to the ecosystem is necessary both for the dynamics and for the complex processes that take place in it [4]. Organic matter is one of the essential components that make up the soil [5][6] and contributes significantly to the assessment of soil conservation and quality [7].
The soil carbon pool is the largest terrestrial pool of organic carbon and plays a vital role in the global carbon cycle and in maintaining ecological balance [8]. Recently, soil management and the great biochar development scale have faced several new challenges [9][10] related to climate change due to increasing temperatures and anthropogenic factors. On the one hand, the continuous degradation of soils because of environmental degradation due to anthropogenic factors and the phenomenon of climate chains are presented as significant issues of environmental concern and are the primary concern of the global community, determining to a considerable extent the performance of productivity and, by extension, the quality, health, and properties of the soil.
On the other hand, increasing population trends, limitations of agricultural land for food production, water scarcity, loss of biodiversity, and maintaining food security in polluted soils [11][12][13] are some components that make the earth unsustainable and make it challenging to function [14]. It is imperative to maintain sustainable management and continuous monitoring of the change in soil properties to increase productivity and minimize or mitigate all those components that negatively affect it. According to the literature review, various types of research demonstrate the positive contribution of biochar in terms of its ability to maintain soil fertility and carbon sequestration for several years [15].
Its properties contribute to the soil’s physical, chemical, and biological properties and plant growth [16]. A sustainable soil resource management strategy is to incorporate biochar into the soil [17], whose use has been known for 2000 years [18], creating an agronomic regime with a positive impact on the environment [19][20][21]. Studies prove its primordial roots and its impressive utilization by the farmers of modern societies [22]. As a soil improver, it is increasingly gaining the trust of its users, which is reflected in many publications [23], and there is an increase in searches of the scientific literature with the word “biochar” [24][25]. Biochar is a porous material whose production is carried out through pyrolysis or hydrothermal treatment of raw biomass [13].
The thermal depolymerization of biomass at high temperatures without the participation of oxygen is called pyrolysis [26][27]. The conceptual approach to the term includes the diverse uses and applications of biochar in various sciences [28]. The use of biochar is an emerging practice for soil amendment that yields promising results [29], attracting worldwide interest, both in research and practical applications [30]. The type of raw material and the production conditions are largely inextricably linked to its physical and chemical properties [31]. These properties determine its physicochemical behavior when applied to the soil and in combination with the properties of the soil [22]. The longevity of the material and its persistence in the environment for a long time are often pointed out in the literature [23]. When it is incorporated into the soil, it is subject to a natural aging process. The natural aging of biochar in the soil is slow and lasts for many years; therefore, accelerated aging techniques are applied as alternative ways of approaching its assessment [32].

2. Properties and Characteristics of Biochar and Factors Affecting Its Performance

Biochar has a high percentage of carbon content and can act as a long-term carbon storage reservoir [33][34], with high stability and large-scale production potential [35][36]. It generally possesses a large surface area, a porous structure, and abundant surface functional groups [37][38]. It contains unstable, stable, volatile, and aromatic compounds due to the presence of carbon.
Rechberger et al. [39] report that biochar is rich in carbon content, and most of it is aromatic in structure and difficult to degrade [40][41]. Hydrogen H, oxygen O, nitrogen N, sulfur S, and ash are present in smaller amounts [18][37][42]. Its structure is highly porous due to carboxyl esterification and aromaticity, with a large specific surface area and low solubility [43]. Biochar properties affected by pyrolysis temperature are surface form and structure [44] and adsorption capacity [45]. Initially, biochar consists of various functional groups, such as hydroxyl (-OH), carboxylate (-COOH), and aldehyde (-CHO), which are located on its surface [18][46][47].
Functional groups vary depending on the type of biochar, the temperature at which it was formed, and the interacting compounds in the soil [48]. During the pyrolysis process, due to increased water loss and dehydration of the biomass, the development of the porous surface of the biochar takes place. Thus, based on pore categorization, biochar is produced with tiny pores (2 nm), medium pores (2–50 nm), and large pores (>50 nm).
Furthermore, the surface area of the biochar plays an essential role because the properties of adsorption and ion exchange are directly related to this property. In essence, it has an internal and an external surface. The first is the walls of the most profound and least open cracks and cavities (micropores). In contrast, the second includes all protrusions, larger cracks, and pores (mesopores and macropores). Macropores contribute less than micropores to the adsorption capacity [49]. Biochar ash is alkaline in nature, and therefore the pH of biochar reflects the ash content [50]. At the same time, its cation exchange capacity is related to its surface, the presence of carboxyl functional groups, the first matter of the biomass used, and the preparation temperature. The stability of the aromatic rings resulting from the increase in temperature turns the biochar into a solid material with high strength for centuries or millennia. Increasing production temperature leads to a decrease in aliphatic structures and the soluble fraction and an increase in the aromatic and insoluble fraction [51].
Therefore, biochar produced at a higher temperature may be more stable than that produced at a lower temperature. Many investigations agree that biochar can be preserved for many years in the soil, even for more than 1000 years [52]. This preservation is attributed to the durability of the biochar carbon and, in particular, to the aromatic carbon, which is more resistant to mineralization under abiotic and biotic stresses.
In essence, C stabilization occurs in aromatic compounds, which certifies the fact of carbon inflow and sequestration in the soil when subjected to a pyrolysis process [53]. In addition, oxygen (O/C) and hydrogen (H/C) content is low, directly related to their durability in the soil. If the pyrolysis temperature is lower than 400 °C (T < 400 °C), the structure of the biochar resembles amorphous carbon. In comparison, if the temperature is higher than 500 °C (T > 500 °C), the structure refers to thermally reduced graphene oxides [54][55]. Research shows that biochar’s pore volume and surface area increase with increasing pyrolysis temperature [56][57].
At the same time, at a temperature >500 °C, the biochar is more resistant to the soil and contains a higher concentration of ash and pH. However, temperatures >700 °C are not acceptable for biochar yield but produce biochar with high stability [10]. Considering the pyrolysis process, which is an exothermic reaction [58], and the thermal behavior of the raw material [59], the control of the pyrolysis temperature can be considered a difficult task. Therefore, the reported temperature may be inaccurate and should be verified. The temperature has a significant impact on the adsorption capacity of biochar [60]. Regarding the effect of pyrolysis on the chemical characteristics of biochar, an increase in pyrolysis temperature also implies an increase in carbon concentration.
However, increasing temperatures gradually decrease H and O concentrations [61]. Generally, at high pyrolysis temperatures (T = 400–700 °C), the feedstock is transformed into stable polycondensed aromatic structures containing high carbon [62]. On the other hand, when the pyrolytic process occurs at a low temperature (400 °C), a biochar product rich in C=O and C-H functional groups is produced. Consequently, the dominant organic compounds are aliphatic or less stable cellulose-like structures at low pyrolysis temperatures, which microbes can quickly degrade. In addition, high pyrolysis temperatures (T > 500 °C) create more hydrophobic biochar, have a larger surface area and pores, have a higher pH due to ash, and are more suitable for organic pollutants.
Low temperatures (500 °C) lead to biochar having a smaller surface area and pores but more oxygen functional groups, making it more suitable for inorganic pollutants [49]. Generally, a higher pyrolysis temperature reduces the amount of solid product that can be made. Still, it improves its quality by giving it a more porous surface, a larger specific surface, more N and K, more C and aromatic substances, and fewer active O groups on its surface [63].
The residence time of biochar during the pyrolysis process significantly affects its porosity and specific surface area [64]. Biochar has a predominantly condensed aromatic structure highly resistant to microbial degradation [65]. The pH is either neutral or alkaline. Alkaline pH sometimes results from high-temperature biomass pyrolysis. However, at lower temperatures, high ionic-strength biochar is produced [65]. According to the literature review, biochar presents multifunctionality that lies in the fact of certain characteristics such as: alkalinity, stability, specific surface area, carbon content, mechanical strength, cation exchange capacity and nutrient retention [66][67][68][69][70][71][72][73][74].

3. The Positive Effect of Biochar on the Ecosystem, with Emphasis on Soil

Aiming at sustainable development, biochar can play an essential role in the long term [18]. Therefore, it is a product gaining more and more research interest, as its use is intended to improve crop yield and soil properties [75]. However, using biochar in soil is not only important for carbon sequestration but also acts as a soil management practice by affecting its physical and chemical properties [17][76]. Soil degradation results in reduced organic matter due to continuous loss, so biochar is a solution for this constant loss of organic matter [33].
In particular, biochar contributes to the assurance and quality of the soil as an environmentally friendly material with a renewable nature without producing pollutants. In addition, its incorporation into the soil could lead to achieving sustainable agriculture [77]. There are three functions of biochar for which its use has been mainly promoted in recent years, which include mitigating climate change [78] by sequestering carbon [79], enhancing the yield of crops, and improving soil structure and functions through multiple beneficial actions. It, thus, plays a vital and active role in agricultural economic development and is an economical solution as an adsorbent material with many environmental applications [80].
One of the significant benefits of biochar is that it helps fight climate change by sequestering carbon dioxide from the atmosphere. It can also be used to rehabilitate problem soils. Biochar has many benefits for agriculture and environmental economics in the long term, so its incorporation into agricultural practices is highly recommended [81]. Biochar is considered a material of global application for the restoration and sustainability of soils due to its physicochemical characteristics and the extensive availability of raw materials [82]. It can promote soil fertility improvement and contribute to climate change mitigation [79]. Biochar as a soil conditioner is gaining more and more interest in research [17]. Three main functions have been discussed in recent years: its contribution to mitigating climate change, its contribution to carbon sequestration in the soil, improving crop yield, and improving soil structure [33][34]. Soil improvement with biochar mainly refers to the improvement of soil nutrient availability (e.g., nitrogen, potassium, and other macronutrients or micronutrients), moisture content, cation exchange capacity (CEC), organic content C, and reducing soil erosion due to improved structure [83]. The effect of biochar on plant growth depends mainly on factors such as the type of biochar, the degree of its incorporation into the soil, the depth of mixing with the soil, the availability of nutrients, the texture of the soil, and the species of cultivated plants [84].
In the last two decades, there has been substantial research interest in studying and further investigating the potential of biochar in terms of increasing productivity, improving soil characteristics, and mitigating climate change. In general, researchers report the beneficial properties of biochar on soil physical, chemical, and biological properties, as well as plant growth [78]. In particular, as a soil conditioner, it can improve long-term physical (bulk density, water holding capacity, permeability, etc.), chemical (nutrient retention, nutrient availability, etc.), and biological properties (microbial population, earthworms, enzyme activity, etc.) for the benefit of plant growth [26]. Consequently, its ability to adsorb pollutants is known. It has already been shown to be widely applicable in wastewater treatment applications, but to date, the mechanisms of pollutant adsorption still need to be fully understood [18]. Additionally, there are many studies on its use as a compound fertilizer due to its high porosity and large surface area [85]. The application of biochar to the environment is not harmful because it has been found to contain several polycyclic aromatic hydrocarbons that are environmentally friendly. However, the commercial use of biochar has not been widely accepted, even though biochar is only a waste [65].

4. Effect of Biochar on Soil’s Physical, Chemical, and Biological Properties

Biochar works in many ways when applied as a technique to improve soil quality. According to the literature review, many reports demonstrate the importance of biochar utilization in the agricultural ecosystem. The positive impact and beneficial effects of biochar incorporation on soil functions have now been demonstrated, as has its effect on soil physical, chemical, and biological properties. However, the factors influencing the long-term behavior of biochar in the environment still need to be better understood. A large body of literature supports the idea that soil amended with biochar has excellent potential to increase crop productivity due to the consequent improvement in soil structure, high nutrient use efficiency (NUE), aeration, porosity, and water holding capacity (WHC) [34]. Biochar exhibits a variety of functions when applied as a technique to improve soil quality. According to the literature review, various reports demonstrate the importance of biochar utilization in the agricultural ecosystem.
The positive impact and beneficial effects of biochar incorporation on soil functions have now been demonstrated, as has its effect on soil’s physical, chemical, and biological properties. However, the factors influencing the long-term fate of biochar in the environment are still poorly understood. An abundant body of literature supports the idea that soil amended with biochar has great potential to increase crop productivity due to the consequent improvement in soil structure, high nutrient use efficiency (NUE), aeration, porosity, and water holding capacity (WHC) [34].
Regarding its effect on soil’s physical properties, biochar contributes to soil texture, grain size, bulk density, aeration, and moisture retention [27]. In addition, an increase in porosity, a decrease in bulk density, and an enhancement of soil aggregation are observed [86][87]. Many studies have shown that biochar affects soil’s water retention and hydrological functions. The results of these studies are variable due to different experimental conditions, including biochar and soil types. Few studies, however, have examined its effect on plant-available water (PAW) and water use efficiency (WUE) [61][88][89][90][91][92][93]. There is an improvement in the movement of water in the soil and an increase in available moisture, mainly in sandy soils with a low content of organic matter, a moderate improvement effect in soils of medium texture, and possibly a reduction in moisture retention for clayey soils. High cation exchange capacity (CEC), high specific surface area, pore volume, and porosity can increase water-holding capacity and improve the effects created by saline soils. Biochar increases soil’s water-holding capacity, especially in arid soils considering climate change.
According to studies, straw biochar formed at 300 °C had a water-holding capacity of 13 × 10−4 ml m−2, but this decreased to 4.1 × 10−4 mL m−2 when the carbonization temperature increased to 700 °C [34][94][95][96][97][98][99][100][101][102][103][104][105][106][107][108].
As a multifactorial carbonaceous and porous material, it improves soil properties and fertility by enhancing the availability of nutrients [109][110][111][112][113][114]. A stoichiometric change in the soil is observed with the simultaneous retention of nutrients [74][115][116]. The enhancement of soil nutrient bioavailability is explained by reducing nutrient leaching and immobilizing toxic elements in contaminated soils [117][118]. In particular, the effect on the potential of nitrogen in the soil is variable since a decrease, an increase, and no change are observed. However, the production feedstock is vital in soil nutrient availability [119][120][121][122].
Biochar is an essential means of long-term increase in organic C [35][102][123][124]. Its incorporation into the soil increases the cation exchange capacity (CEC) and the degree of base saturation, thereby increasing the retention of nutrients [100][125]. According to research, the improvement of CEC reached up to 20% and electrical conductivity (EC) up to 124.6% [34][100][126][127][128][129][130], increasing soil pH and reducing soil acidity by 31.9%. Significant crop production and nutritional value increases are observed when its application is in acidic and poor soils. Conversely, it can have controversial results in alkaline and fertile soils, causing a neutral or negative effect on plant tissue production [68][89][127][131][132][133][134]. In terms of its contribution to soil’s biological properties shows growth in microbial activity and diversity [135][136][137][138]. It immobilizes the microbial nitrogen (N) of the biomass and stabilizes the microbial nitrogen (N) of the soil [128][139][140].
Its positive effects appear in the reduction in fertilizer leaching and improvement of soil aeration [100][141], in the improvement of stream flow resistance [142], in the decontamination of air and water soil [143][144][145][146], and in the general improvement of soil aeration and hydrological soil functions [147][148][149]. The effect of biochar on plant growth and crop yield depends on the crop type, soil type, biochar properties, rate, method, and frequency of its application. Crop yield responses are generally reported as positive. Biochar application has been reported to increase plant productivity by approximately 10% and up to 25% for aboveground biomass. Other research investigates biochar’s effect on increasing corn’s biomass due to increased soil pH and CEC. Increase grain and maize yields by 150% and 98% with 15 and 20 t/ha biochar, respectively, to improve soil’s physical and chemical properties.
The effectiveness of biochar in improving plant productivity is variable, considering climate change, soil properties, crops, and experimental conditions. These differences could also be explained by the different pyrolysis processes and soil interactions with biotic and abiotic factors [68][150][151][152][153][154]. A general increase in crop productivity and yield is observed. Biochar showed an increase in available plant water content of 33–45% in coarse-grained soils and 9–14% in clay soils. A tremendous increase was achieved by adding 30–70 Mg/ha. At the same time, there was an average increase of 27% in the rate of photosynthesis in C3 plants but no effect in C4 plants.
Increased stomatal conductance, transpiration rate, and chlorophyll content were attributed to the combined effects of biochar, water availability, and nitrogen fertilization, sequestering soil carbon as a long-term carbon pool and reducing emissions of greenhouse gases (GHG) such as carbon dioxide (CO2), nitrous oxide (N2O), and methane (CH4). However, the accuracy of its application remains controversial. Biochar application to paddy fields caused a 12% increase in CO2 emissions but a decrease in NO2 emissions. In a pasture ecosystem, biochar application showed no change in CO2 or NO2 emissions. Other research showed a CH4 reduction in soybean crops after biochar addition due to increased soil pH and strong adsorption capacity and a 50% N2O reduction in soybean plots on acidic soils. With the addition of biochar, there is a spatial reorganization of C within soil particles, but the mechanisms remain unclear [101][125][148][155][156][157][158][159][160][161][162].

Mitigating Climate Change by Removing CO2 from the Atmosphere

Biochar’s proposed climate change mitigation mechanisms are its molecular structure, dominated by aromatic carbons that make it more resistant to microbial decomposition, allowing it to remain in the soil for thousands of years and potentially limiting greenhouse gas emissions. The change in CO2 emission rate can be related to soil temperature, soil type, pyrolysis temperature, and biochar application rate [34][35][36][140][163][164][165][166][167][168][169][170][171][172][173][174].
Biochar provides a large capacity to adsorb heavy metals from the soil due to its large surface area, multi-layered porous structure, abundant surface functional groups, increased soil redox potential, pH, organic matter content, and cation exchange capacity [97][175][176]. At the same time, it shows a reduction in the leaching of nutrients into the soil due to the high CEC of biochar [177]. The modern literature references the involvement of biochar in other areas as well. The use of biochar in the horticulture industry for fruit and vegetable production is considered [159][178]. In addition, it contributes to the restoration of forest ecosystems and urban soils due to its ability to adsorb and immobilize organic substances [179]. However, it exhibits combinatorial and synergistic action with other materials such as compost, manure, paper mill sludge, and biosolids [180][181][182].

References

  1. Pepper, I.L.; Gerba, C.P.; Newby, D.T.; Rice, C.W. Soil: A Public Health Threat or Savior? Crit. Rev. Environ. Sci. Technol. 2009, 39, 416–432.
  2. Liu, Q.; Yang, D.; Cao, L.; Anderson, B. Assessment and Prediction of Carbon Storage Based on Land Use/Land Cover Dynamics in the Tropics: A Case Study of Hainan Island, China. Land 2022, 11, 244.
  3. Sabetizade, M.; Gorji, M.; Roudier, P.; Zolfaghari, A.A.; Keshavarzi, A. Combination of MIR spectroscopy and environmental covariates to predict soil organic carbon in a semi-arid region. Catena 2020, 196, 104844.
  4. He, M.; Xiong, X.; Wang, L.; Hou, D.; Bolan, N.S.; Ok, Y.S.; Rinklebe, J.; Tsang, D.C. A critical review on performance indicators for evaluating soil biota and soil health of biochar-amended soils. J. Hazard. Mater. 2021, 414, 125378.
  5. Di Palma, L.; Ferrantelli, P.; Merli, C.; Petrucci, E.; Pitzolu, I. Influence of Soil Organic Matter on Copper Extraction from Contaminated Soil. Soil Sediment Contam. Int. J. 2007, 16, 323–335.
  6. Cuffney, T.F.; Brightbill, R.A.; May, J.T.; Waite, I.R. Responses of benthic macroinvertebrates to environmental changes associated with urbanization in nine metropolitan areas. Ecol. Appl. A Publ. Ecol. Soc. Am. 2010, 20, 1384–1401.
  7. Kim, M.-J.; Lee, H.-I.; Choi, J.-H.; Lim, K.J.; Mo, C. Development of a Soil Organic Matter Content Prediction Model Based on Supervised Learning Using Vis-NIR/SWIR Spectroscopy. Sensors 2022, 22, 5129.
  8. Zhu, C.; Wei, Y.; Zhu, F.; Lu, W.; Fang, Z.; Li, Z.; Pan, J. Digital Mapping of Soil Organic Carbon Based on Machine Learning and Regression Kriging. Sensors 2022, 22, 8997.
  9. Thomas, F.; Petzold, R.; Landmark, S.; Mollenhauer, H.; Becker, C.; Werban, U. Estimating Forest Soil Properties for Humus Assessment—Is Vis-NIR the Way to Go? Remote Sens. 2022, 14, 1368.
  10. Osman, A.I.; Fawzy, S.; Farghali, M.; El-Azazy, M.; Elgarahy, A.M.; Fahim, R.A.; Maksoud, M.I.A.A.; Ajlan, A.A.; Yousry, M.; Saleem, Y.; et al. Biochar for agronomy, animal farming, anaerobic digestion, composting, water treatment, soil remediation, construction, energy storage, and carbon sequestration: A review. Environ. Chem. Lett. 2022, 20, 2385–2485.
  11. Kumar, A.; Bhattacharya, T.; Mukherjee, S.; Sarkar, B. A perspective on biochar for repairing damages in the soil–plant system caused by climate change-driven extreme weather events. Biochar 2022, 4, 22.
  12. De la Rosa, J.M.; Campos, P.; Diaz-Espejo, A. Soil Biochar Application: Assessment of the Effects on Soil Water Properties, Plant Physiological Status, and Yield of Super-Intensive Olive Groves under Controlled Irrigation Conditions. Agronomy 2022, 12, 2321.
  13. Wang, L.; O’connor, D.; Rinklebe, J.; Ok, Y.S.; Tsang, D.C.; Shen, Z.; Hou, D. Biochar Aging: Mechanisms, Physicochemical Changes, Assessment, And Implications for Field Applications. Environ. Sci. Technol. 2020, 54, 14797–14814.
  14. Tsakiridis, N.L.; Theocharis, J.B.; Symeonidis, A.L.; Zalidis, G.C. Improving the predictions of soil properties from VNIR–SWIR spectra in an unlabeled region using semi-supervised and active learning. Geoderma 2021, 387, 114830.
  15. Moiwo, J.P.; Wahab, A.; Kangoma, E.; Blango, M.M.; Ngegba, M.P.; Suluku, R. Effect of Biochar Application Depth on Crop Productivity Under Tropical Rainfed Conditions. Appl. Sci. 2019, 9, 2602.
  16. Joseph, S.; Cowie, A.L.; Van Zwieten, L.; Bolan, N.; Budai, A.; Buss, W.; Cayuela, M.L.; Graber, E.R.; Ippolito, J.A.; Kuzyakov, Y.; et al. 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.
  17. Chen, W.; Meng, J.; Han, X.; Lan, Y.; Zhang, W. Past, present, and future of biochar. Biochar 2019, 1, 75–87.
  18. Wikurendra, E.A. Biochar: A Review of its History, Characteristics, Factors that Influence its Yield, Methods of Production, Application in Wastewater Treatment and Recent Development. Biointerface Res. Appl. Chem. 2021, 12, 7914–7926.
  19. Fang, Y.; Singh, B.P.; Krull, E. Biochar carbon stability in four contrasting soils. Eur. J. Soil Sci. 2013, 65, 60–71.
  20. Whitman, T.; Lehmann, J. Biochar—One way forward for soil carbon in offset mechanisms in Africa? Environ. Sci. Policy 2009, 12, 1024–1027.
  21. Zimmerman, A.R. Abiotic and Microbial Oxidation of Laboratory-Produced Black Carbon (Biochar). Environ. Sci. Technol. 2010, 44, 1295–1301.
  22. Tan, S.; Narayanan, M.; Huong, D.T.T.; Ito, N.; Unpaprom, Y.; Pugazhendhi, A.; Chi, N.T.L.; Liu, J. A perspective on the interaction between biochar and soil microbes: A way to regain soil eminence. Environ. Res. 2022, 214, 113832.
  23. Bednik, M.; Medyńska-Juraszek, A.; Ćwieląg-Piasecka, I. Effect of Six Different Feedstocks on Biochar’s Properties and Expected Stability. Agronomy 2022, 12, 1525.
  24. Torabian, S.; Qin, R.; Noulas, C.; Lu, Y.; Wang, G. Biochar: An organic amendment to crops and an environmental solution. AIMS Agric. Food 2021, 6, 401–415.
  25. Sachdeva, S.; Kumar, R.; Sahoo, P.K.; Nadda, A.K. Recent advances in biochar amendments for immobilization of heavy metals in an agricultural ecosystem: A systematic review. Environ. Pollut. 2023, 319, 120937.
  26. Layek, J.; Narzari, R.; Hazarika, S.; Das, A.; Rangappa, K.; Devi, S.; Balusamy, A.; Saha, S.; Mandal, S.; Idapuganti, R.G.; et al. Prospects of Biochar for Sustainable Agriculture and Carbon Sequestration: An Overview for Eastern Himalayas. Sustainability 2022, 14, 6684.
  27. 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.
  28. Fellet, G.; Conte, P.; Bortolotti, V.; Zama, F.; Landi, G.; Martino, D.F.C.; Ferro, V.; Marchiol, L.; Meo, P.L. Changes in Physicochemical Properties of Biochar after Addition to Soil. Agriculture 2022, 12, 320.
  29. Brassard, P.; Godbout, S.; Lévesque, V.; Palacios, J.H.; Raghavan, V.; Ahmed, A.; Hogue, R.; Jeanne, T.; Verma, M. Biochar for soil amendment. Prod. Charact. Appl. 2019, 109–146.
  30. Li, X.; Wang, C.; Zhang, J.; Liu, J.; Liu, B.; Chen, G. Preparation and application of magnetic biochar in water treatment: A critical review. Sci. Total. Environ. 2019, 711, 134847.
  31. Yuan, J.-H.; Xu, R.-K.; Zhang, H. The forms of alkalis in the biochar produced from crop residues at different temperatures. Bioresour. Technol. 2011, 102, 3488–3497.
  32. Spokas, K.A.; Novak, J.M.; Masiello, C.A.; Johnson, M.G.; Colosky, E.C.; Ippolito, J.A.; Trigo, C. Physical Disintegration of Biochar: An Overlooked Process. Environ. Sci. Technol. Lett. 2014, 1, 326–332.
  33. Jones, D.L.; Rousk, J.; Edwards-Jones, G.; DeLuca, T.H.; Murphy, D.V. Biochar-mediated changes in soil quality and plant growth in a three year field trial. Soil Biol. Biochem. 2012, 45, 113–124.
  34. Alkharabsheh, H.M.; Seleiman, M.F.; Battaglia, M.L.; Shami, A.; Jalal, R.S.; Alhammad, B.A.; Almutairi, K.F.; Al–Saif, A.M. Biochar and Its Broad Impacts in Soil Quality and Fertility, Nutrient Leaching and Crop Productivity: A Review. Agronomy 2021, 11, 993.
  35. Lehmann, J.; Gaunt, J.; Rondon, M. Bio-char Sequestration in Terrestrial Ecosystems—A Review. Mitig. Adapt. Strat. Glob. Chang. 2006, 11, 403–427.
  36. Gross, A.; Bromm, T.; Glaser, B. Soil Organic Carbon Sequestration after Biochar Application: A Global Meta-Analysis. Agronomy 2021, 11, 2474.
  37. Tan, X.; Liu, Y.; Zeng, G.; Wang, X.; Hu, X.; Gu, Y.; Yang, Z. Application of biochar for the removal of pollutants from aqueous solutions. Chemosphere 2015, 125, 70–85.
  38. Xu, R.-K.; Xiao, S.-C.; Yuan, J.-H.; Zhao, A.-Z. Adsorption of methyl violet from aqueous solutions by the biochars derived from crop residues. Bioresour. Technol. 2011, 102, 10293–10298.
  39. Rechberger, M.V.; Kloss, S.; Wang, S.-L.; Lehmann, J.; Rennhofer, H.; Ottner, F.; Wriessnig, K.; Daudin, G.; Lichtenegger, H.; Soja, G.; et al. Enhanced Cu and Cd sorption after soil aging of woodchip-derived biochar: What were the driving factors? Chemosphere 2018, 216, 463–471.
  40. Singh, B.P.; Cowie, A.L.; Smernik, R.J. Biochar Carbon Stability in a Clayey Soil As a Function of Feedstock and Pyrolysis Temperature. Environ. Sci. Technol. 2012, 46, 11770–11778.
  41. An, N.; Zhang, L.; Liu, Y.; Shen, S.; Li, N.; Wu, Z.; Yang, J.; Han, W.; Han, X. Biochar application with reduced chemical fertilizers improves soil pore structure and rice productivity. Chemosphere 2022, 298, 134304.
  42. 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.
  43. Ahmadi, A.; Emami, M.; Daccache, A.; He, L. Soil Properties Prediction for Precision Agriculture Using Visible and Near-Infrared Spectroscopy: A Systematic Review and Meta-Analysis. Agronomy 2021, 11, 433.
  44. Ahmad, M.; Rajapaksha, A.U.; Lim, J.E.; Zhang, M.; Bolan, N.; Mohan, D.; Vithanage, M.; Lee, S.S.; Ok, Y.S. Biochar as a sorbent for contaminant management in soil and water: A review. Chemosphere 2014, 99, 19–33.
  45. Liu, Z.; Zhu, M.; Wang, J.; Liu, X.; Guo, W.; Zheng, J.; Bian, R.; Wang, G.; Zhang, X.; Cheng, K.; et al. The responses of soil organic carbon mineralization and microbial communities to fresh and aged biochar soil amendments. GCB Bioenergy 2019, 11, 1408–1420.
  46. Fan, Q.; Sun, J.; Chu, L.; Cui, L.; Quan, G.; Yan, J.; Hussain, Q.; Iqbal, M. Effects of chemical oxidation on surface oxygen-containing functional groups and adsorption behavior of biochar. Chemosphere 2018, 207, 33–40.
  47. Lehmann, J. Bio-energy in the black. Front. Ecol. Environ. 2007, 5, 381–387.
  48. 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.
  49. Amalina, F.; Krishnan, S.; Zularisam, A.; Nasrullah, M. Biochar and sustainable environmental development towards adsorptive removal of pollutants: Modern advancements and future insight. Process. Saf. Environ. Prot. 2023, 173, 715–728.
  50. Jiang, B.; Lin, Y.; Mbog, J.C. Biochar derived from swine manure digestate and applied on the removals of heavy metals and antibiotics. Bioresour. Technol. 2018, 270, 603–611.
  51. Yang, Y.; Sun, K.; Han, L.; Chen, Y.; Liu, J.; Xing, B. Biochar stability and impact on soil organic carbon mineralization depend on biochar processing, aging and soil clay content. Soil Biol. Biochem. 2022, 169, 108657.
  52. Kuzyakov, Y.; Subbotina, I.; Chen, H.; Bogomolova, I.; Xu, X. Black carbon decomposition and incorporation into soil microbial biomass estimated by 14C labeling. Soil Biol. Biochem. 2009, 41, 210–219.
  53. Tozzi, F.V.D.N.; Coscione, A.R.; Puga, A.P.; Carvalho, C.S.; Cerri, C.E.P.; Andrade, C. Carbon stability and biochar aging process after soil application. Hortic. Int. J. 2019, 3, 320–329.
  54. Brodowski, S.; John, B.; Flessa, H.; Amelung, W. Aggregate-occluded black carbon in soil. Eur. J. Soil Sci. 2006, 57, 539–546.
  55. McDonald-Wharry, J.; Manley-Harris, M.; Pickering, K. Carbonisation of biomass-derived chars and the thermal reduction of a graphene oxide sample studied using Raman spectroscopy. Carbon 2013, 59, 383–405.
  56. Weber, K.; Quicker, P. Properties of biochar. Fuel 2018, 217, 240–261.
  57. Fu, P.; Hu, S.; Xiang, J.; Sun, L.; Su, S.; Wang, J. Evaluation of the porous structure development of chars from pyrolysis of rice straw: Effects of pyrolysis temperature and heating rate. J. Anal. Appl. Pyrolysis 2012, 98, 177–183.
  58. Ripberger, G.D.; Jones, J.R.; Paterson, A.H.J.; Holt, R. Is It Possible to Produce Biochar at Different Highest Treatment Temperatures in the Pyrolysis Range?—The Exothermic Nature of Pyrolysis. In Asia Pacific Confederation of Chemical Engineering Congress 2015: APCChE 2015, Incorporating CHEMECA 2015; Engineers Australia: Melbourne, Australia, 2015; pp. 1950–1957.
  59. Di Blasi, C.; Branca, C.; Sarnataro, F.E.; Gallo, A. Thermal Runaway in the Pyrolysis of Some Lignocellulosic Biomasses. Energy Fuels 2014, 28, 2684–2696.
  60. Ahmad, A.; Khan, N.; Giri, B.S.; Chowdhary, P.; Chaturvedi, P. Removal of methylene blue dye using rice husk, cow dung and sludge biochar: Characterization, application, and kinetic studies. Bioresour. Technol. 2020, 306, 123202.
  61. Novak, J.M.; Lima, I.; Xing, B.; Gaskin, J.W.; Steiner, C.; Das, K.C.; Ahmedna, M.; Rehrah, D.; Watts, D.W.; Busscher, W.J.; et al. Characterization of Designer Biochar Produced at Different Temperatures and Their Effects on a Loamy Sand. Ann. Environ. Sci. 2009, 3. Available online: https://openjournals.neu.edu/aes/journal/article/view/v3art5 (accessed on 2 June 2023).
  62. Kaal, J.; Cortizas, A.M.; Reyes, O.; Soliño, M. Molecular characterization of Ulex europaeus biochar obtained from laboratory heat treatment experiments—A pyrolysis–GC/MS study. J. Anal. Appl. Pyrolysis 2012, 95, 205–212.
  63. Aller, M.F. Biochar properties: Transport, fate, and impact. Crit. Rev. Environ. Sci. Technol. 2016, 46, 1183–1296.
  64. Brassard, P.; Godbout, S.; Raghavan, V. Soil biochar amendment as a climate change mitigation tool: Key parameters and mechanisms involved. J. Environ. Manag. 2016, 181, 484–497.
  65. Okareh, O.T.; Gbadebo, A.O. Enhancement of Soil Health Using Biochar. In Applications of Biochar for Environmental Safety; IntechOpen: London, UK, 2020.
  66. Chintala, R.; Mollinedo, J.; Schumacher, T.E.; Malo, D.D.; Julson, J.L. Effect of biochar on chemical properties of acidic soil. Arch. Agron. Soil Sci. 2013, 60, 393–404.
  67. Mimmo, T.; Panzacchi, P.; Baratieri, M.; Davies, C.; Tonon, G. Effect of pyrolysis temperature on miscanthus (Miscanthus × giganteus) biochar physical, chemical and functional properties. Biomass-Bioenergy 2014, 62, 149–157.
  68. Biederman, L.A.; Harpole, W.S. Biochar and its effects on plant productivity and nutrient cycling: A meta-analysis. GCB Bioenergy 2013, 5, 202–214.
  69. Suliman, W.; Harsh, J.B.; Abu-Lail, N.I.; Fortuna, A.-M.; Dallmeyer, I.; Garcia-Pérez, M. The role of biochar porosity and surface functionality in augmenting hydrologic properties of a sandy soil. Sci. Total. Environ. 2017, 574, 139–147.
  70. 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.
  71. Syuhada, A.B.; Shamshuddin, J.; Fauziah, C.I.; Rosenani, A.B.; Arifin, A. Biochar as soil amendment: Impact on chemical properties and corn nutrient uptake in a Podzol. Can. J. Soil Sci. 2016, 96, 400–412.
  72. Agegnehu, G.; Bass, A.M.; Nelson, P.N.; Bird, M.I. Benefits of biochar, compost and biochar–compost for soil quality, maize yield and greenhouse gas emissions in a tropical agricultural soil. Sci. Total. Environ. 2016, 543, 295–306.
  73. Beusch, C.; Cierjacks, A.; Böhm, J.; Mertens, J.; Bischoff, W.-A.; Filho, J.C.D.A.; Kaupenjohann, M. Biochar vs. clay: Comparison of their effects on nutrient retention of a tropical Arenosol. Geoderma 2019, 337, 524–535.
  74. Hui, D. Effects of Biochar Application on Soil Properties, Plant Biomass Production, and Soil Greenhouse Gas Emissions: A Mini-Review. Agric. Sci. 2021, 12, 213–236.
  75. Gaskin, J.W.; Steiner, C.; Harris, K.; Das, K.C.; Bibens, B. Effect of Low-Temperature Pyrolysis Conditions on Biochar for Agricultural Use. Trans. ASABE 2008, 51, 2061–2069.
  76. Ye, S.; Zeng, G.; Wu, H.; Liang, J.; Zhang, C.; Dai, J.; Xiong, W.; Song, B.; Wu, S.; Yu, J. The effects of activated biochar addition on remediation efficiency of co-composting with contaminated wetland soil. Resour. Conserv. Recycl. 2018, 140, 278–285.
  77. Ayaz, M.; Feizienė, D.; Tilvikienė, V.; Akhtar, K.; Stulpinaitė, U.; Iqbal, R. Biochar Role in the Sustainability of Agriculture and Environment. Sustainability 2021, 13, 1330.
  78. Snyder, H. Literature review as a research methodology: An overview and guidelines. J. Bus. Res. 2019, 104, 333–339.
  79. Agarwal, H.; Kashyap, V.H.; Mishra, A.; Bordoloi, S.; Singh, P.K.; Joshi, N.C. Biochar-based fertilizers and their applications in plant growth promotion and protection. 3 Biotech 2022, 12, 136.
  80. Zhou, Y.; Qin, S.; Verma, S.; Sar, T.; Sarsaiya, S.; Ravindran, B.; Liu, T.; Sindhu, R.; Patel, A.K.; Binod, P.; et al. Production and beneficial impact of biochar for environmental application: A comprehensive review. Bioresour. Technol. 2021, 337, 125451.
  81. Rehman, H.A.; Razzaq, R. Benefits of Biochar on the Agriculture and Environment—A Review. J. Environ. Anal. Chem. 2017, 4, 3.
  82. Qiu, M.; Liu, L.; Ling, Q.; Cai, Y.; Yu, S.; Wang, S.; Fu, D.; Hu, B.; Wang, X. Biochar for the removal of contaminants from soil and water: A review. Biochar 2022, 4, 19.
  83. Nath, H.; Sarkar, B.; Mitra, S.; Bhaladhare, S. Biochar from Biomass: A Review on Biochar Preparation Its Modification and Impact on Soil Including Soil Microbiology. Geomicrobiol. J. 2022, 39, 373–388.
  84. O’Connor, D.; Peng, T.; Zhang, J.; Tsang, D.C.; Alessi, D.S.; Shen, Z.; Bolan, N.S.; Hou, D. Biochar application for the remediation of heavy metal polluted land: A review of in situ field trials. Sci. Total Environ. 2018, 619–620, 815–826.
  85. Mullai, P.; Vishali, S.; Kobika, P.; Dhivya, K.S.; Mukund, A.; Sriraaman, M. Biochar Production and Its Basket Full Of Benefits—A Review. ECS Trans. 2022, 107, 18747–18752.
  86. Ding, Y.; Liu, Y.; Liu, S.; Li, Z.; Tan, X.; Huang, X.; Zeng, G.; Zhou, L.; Zheng, B. Biochar to improve soil fertility. A review. Agron. Sustain. Dev. 2016, 36, 36.
  87. Baiamonte, G.; Crescimanno, G.; Parrino, F.; De Pasquale, C. Effect of biochar on the physical and structural properties of a sandy soil. Catena 2019, 175, 294–303.
  88. 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.
  89. Artiola, J.F.; Rasmussen, C.; Freitas, R. Effects of a Biochar-Amended Alkaline Soil on the Growth of Romaine Lettuce and Bermudagrass. Soil Sci. 2012, 177, 561–570.
  90. Basso, A.S.; Miguez, F.E.; Laird, D.A.; Horton, R.; Westgate, M. Assessing potential of biochar for increasing water-holding capacity of sandy soils. GCB Bioenergy 2012, 5, 132–143.
  91. Abel, S.; Peters, A.; Trinks, S.; Schonsky, H.; Facklam, M.; Wessolek, G. Impact of biochar and hydrochar addition on water retention and water repellency of sandy soil. Geoderma 2013, 202–203, 183–191.
  92. Rogovska, N.; Laird, D.; Cruse, R.M.; Trabue, S.; Heaton, E. Germination Tests for Assessing Biochar Quality. J. Environ. Qual. 2012, 41, 1014–1022.
  93. Ma, N.; Zhang, L.; Zhang, Y.; Yang, L.; Yu, C.; Yin, G.; Doane, T.A.; Wu, Z.; Zhu, P.; Ma, X. Biochar Improves Soil Aggregate Stability and Water Availability in a Mollisol after Three Years of Field Application. PLoS ONE 2016, 11, e0154091.
  94. El-Naggar, A.; Lee, S.S.; Rinklebe, J.; Farooq, M.; Song, H.; Sarmah, A.K.; Zimmerman, A.R.; Ahmad, M.; Shaheen, S.M.; Ok, Y.S. Biochar application to low fertility soils: A review of current status, and future prospects. Geoderma 2019, 337, 536–554.
  95. Lim, T.; Spokas, K.; Feyereisen, G.; Novak, J. Predicting the impact of biochar additions on soil hydraulic properties. Chemosphere 2016, 142, 136–144.
  96. Novak, J.M.; Busscher, W.J.; Watts, D.W.; Amonette, J.E.; Ippolito, J.A.; Lima, I.M.; Gaskin, J.; Das, K.C.; Steiner, C.; Ahmedna, M.; et al. Biochars Impact on Soil-Moisture Storage in an Ultisol and Two Aridisols. Soil Sci. 2012, 177, 310–320.
  97. 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.
  98. Sun, F.; Lu, S. Biochars improve aggregate stability, water retention, and pore-space properties of clayey soil. J. Plant Nutr. Soil Sci. 2013, 177, 26–33.
  99. Liu, P.; Liu, W.-J.; Jiang, H.; Chen, J.-J.; Li, W.-W.; Yu, H.-Q. Modification of bio-char derived from fast pyrolysis of biomass and its application in removal of tetracycline from aqueous solution. Bioresour. Technol. 2012, 121, 235–240.
  100. Laird, D.A.; Fleming, P.; Davis, D.D.; Horton, R.; Wang, B.; Karlen, D.L. Impact of biochar amendments on the quality of a typical Midwestern agricultural soil. Geoderma 2010, 158, 443–449.
  101. Rogovska, N.; Laird, D.A.; Rathke, S.J.; Karlen, D.L. Biochar impact on Midwestern Mollisols and maize nutrient availability. Geoderma 2014, 230–231, 340–347.
  102. 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.
  103. Atkinson, C.J.; Fitzgerald, J.D.; Hipps, N.A. Potential mechanisms for achieving agricultural benefits from biochar application to temperate soils: A review. Plant Soil 2010, 337, 1–18.
  104. Busscher, W.J.; Novak, J.M.; Evans, D.E.; Watts, D.W.; Niandou, M.A.S.; Ahmedna, M. Influence of Pecan Biochar on Physical Properties of a Norfolk Loamy Sand. Soil Sci. 2010, 175, 10–14.
  105. Głuszek, S.; Sas-Paszt, L.; Sumorok, B.; Kozera, R. Biochar-Rhizosphere Interactions—A Review. Pol. J. Microbiol. 2017, 66, 151–161.
  106. Arif, M.; Ilyas, M.; Riaz, M.; Ali, K.; Shah, K.; Haq, I.U.; Fahad, S. Biochar improves phosphorus use efficiency of organic-inorganic fertilizers, maize-wheat productivity and soil quality in a low fertility alkaline soil. Field Crop. Res. 2017, 214, 25–37.
  107. Ojeda, G.; Mattana, S.; Àvila, A.; Alcañiz, J.M.; Volkmann, M.; Bachmann, J. Are soil–water functions affected by biochar application? Geoderma 2015, 249–250, 1–11.
  108. Kammann, C.I.; Schmidt, H.-P.; Messerschmidt, N.; Linsel, S.; Steffens, D.; Müller, C.; Koyro, H.-W.; Conte, P.; Joseph, S. Plant growth improvement mediated by nitrate capture in co-composted biochar. Sci. Rep. 2015, 5, 11080.
  109. Baiamonte, G.; De Pasquale, C.; Marsala, V.; Cimò, G.; Alonzo, G.; Crescimanno, G.; Conte, P. Structure alteration of a sandy-clay soil by biochar amendments. J. Soils Sediments 2015, 15, 816–824.
  110. El Barnossi, A.; Moussaid, F.; Housseini, A.I. Tangerine, banana and pomegranate peels valorisation for sustainable environment: A review. Biotechnol. Rep. 2020, 29, e00574.
  111. Gagnon, B.; Ziadi, N.; Manirakiza, E. Co-application of wood biochar and paper mill biosolids affects yield and short-term nitrogen and phosphorus availability in temperate loamy soils. Can. J. Soil Sci. 2022, 102, 131–146.
  112. Abedin, J.; Unc, A. The utility of biochar for increasing the fertility of new agricultural lands converted from boreal forests. Can. J. Soil Sci. 2022, 102, 165–176.
  113. Lévesque, V.; Gagnon, B.; Ziadi, N. Soil Mehlich-3-extractable elements as affected by the addition of biochars to a clay soil co-amended with or without a compost. Can. J. Soil Sci. 2022, 102, 97–107.
  114. Prendergast-Miller, M.T.; Duvall, M.; Sohi, S.P. Biochar-root interactions are mediated by biochar nutrient content and impacts on soil nutrient availability. Eur. J. Soil Sci. 2013, 65, 173–185.
  115. 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.
  116. 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.
  117. Igalavithana, A.D.; Kwon, E.E.; Vithanage, M.; Rinklebe, J.; Moon, D.H.; Meers, E.; Tsang, D.C.; Ok, Y.S. Soil lead immobilization by biochars in short-term laboratory incubation studies. Environ. Int. 2019, 127, 190–198.
  118. Chen, P.; Liu, Y.; Mo, C.; Jiang, Z.; Yang, J.; Lin, J. Microbial mechanism of biochar addition on nitrogen leaching and retention in tea soils from different plantation ages. Sci. Total. Environ. 2020, 757, 143817.
  119. Lentz, R.D.; Ippolito, J.A.; Spokas, K.A. Biochar and Manure Effects on Net Nitrogen Mineralization and Greenhouse Gas Emissions from Calcareous Soil under Corn. Soil Sci. Soc. Am. J. 2014, 78, 1641–1655.
  120. Domene, X.; Mattana, S.; Hanley, K.; Enders, A.; Lehmann, J. Medium-term effects of corn biochar addition on soil biota activities and functions in a temperate soil cropped to corn. Soil Biol. Biochem. 2014, 72, 152–162.
  121. Gaskin, J.W.; Speir, R.A.; Harris, K.; Das, K.C.; Lee, R.D.; Morris, L.A.; Fisher, D.S. Effect of Peanut Hull and Pine Chip Biochar on Soil Nutrients, Corn Nutrient Status, and Yield. Agron. J. 2010, 102, 623–633.
  122. Mia, S.; van Groenigen, J.; van de Voorde, T.; Oram, N.; Bezemer, T.; Mommer, L.; Jeffery, S. Biochar application rate affects biological nitrogen fixation in red clover conditional on potassium availability. Agric. Ecosyst. Environ. 2014, 191, 83–91.
  123. Lehmann, J.; Joseph, S. Biochar for Environmental Management: Science, Technology and Implementation; Routledge: Oxfordshire, UK, 2015.
  124. Spokas, K.A.; Cantrell, K.B.; Novak, J.M.; Archer, D.W.; Ippolito, J.A.; Collins, H.P.; Boateng, A.A.; Lima, I.M.; Lamb, M.C.; McAloon, A.J.; et al. Biochar: A Synthesis of Its Agronomic Impact beyond Carbon Sequestration. J. Environ. Qual. 2012, 41, 973–989.
  125. Zhang, A.; Bian, R.; Pan, G.; Cui, L.; Hussain, Q.; Li, L.; Zheng, J.; Zheng, J.; Zhang, X.; Han, X.; et al. Effects of biochar amendment on soil quality, crop yield and greenhouse gas emission in a Chinese rice paddy: A field study of 2 consecutive rice growing cycles. Field Crop. Res. 2012, 127, 153–160.
  126. Saha, A.; Basak, B.B. Scope of value addition and utilization of residual biomass from medicinal and aromatic plants. Ind. Crop. Prod. 2020, 145, 111979.
  127. Oguntunde, P.G.; Fosu, M.; Ajayi, A.E.; van de Giesen, N. Effects of charcoal production on maize yield, chemical properties and texture of soil. Biol. Fertil. Soils 2004, 39, 295–299.
  128. Foster, E.J.; Hansen, N.; Wallenstein, M.; Cotrufo, M.F. Biochar and manure amendments impact soil nutrients and microbial enzymatic activities in a semi-arid irrigated maize cropping system. Agric. Ecosyst. Environ. 2016, 233, 404–414.
  129. DeLuca, T.; Fajvan, M.A.; Miller, G. Diameter-Limit Harvesting: Effects of Residual Trees on Regeneration Dynamics in Appalachian Hardwoods. North. J. Appl. For. 2009, 26, 52–60.
  130. Yuan, Y.; Bolan, N.; Prévoteau, A.; Vithanage, M.; Biswas, J.K.; Ok, Y.S.; Wang, H. Applications of biochar in redox-mediated reactions. Bioresour. Technol. 2017, 246, 271–281.
  131. Van Zwieten, L.; Kimber, S.; Morris, S.; Chan, K.Y.; Downie, A.; Rust, J.; Joseph, S.; Cowie, A. Effects of biochar from slow pyrolysis of papermill waste on agronomic performance and soil fertility. Plant Soil 2009, 327, 235–246.
  132. 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.
  133. Marks, E.A.N.; Alcañiz, J.M.; Domene, X. Unintended effects of biochars on short-term plant growth in a calcareous soil. Plant Soil 2014, 385, 87–105.
  134. Upadhyay, K.P.; George, D.; Swift, R.S.; Galea, V. The Influence of Biochar on Growth of Lettuce and Potato. J. Integr. Agric. 2014, 13, 541–546.
  135. Jones, D.; Magthab, E.; Gleeson, D.; Hill, P.; Sánchez-Rodríguez, A.; Roberts, P.; Ge, T.; Murphy, D. Microbial competition for nitrogen and carbon is as intense in the subsoil as in the topsoil. Soil Biol. Biochem. 2018, 117, 72–82.
  136. Liu, S.; Zhang, Y.; Zong, Y.; Hu, Z.; Wu, S.; Zhou, J.; Jin, Y.; Zou, J. Response of soil carbon dioxide fluxes, soil organic carbon and microbial biomass carbon to biochar amendment: A meta-analysis. GCB Bioenergy 2016, 8, 392–406.
  137. Xu, H.-J.; Wang, X.-H.; Li, H.; Yao, H.-Y.; Su, J.-Q.; Zhu, Y.-G. Biochar Impacts Soil Microbial Community Composition and Nitrogen Cycling in an Acidic Soil Planted with Rape. Environ. Sci. Technol. 2014, 48, 9391–9399.
  138. Thies, E.; Rilling, M. Characteristics of biochar: Biological properties. In Biochar for Environmental Management: Science and Technology; Lehmann, J., Joseph, S., Eds.; Earthscan: London, UK, 2009; pp. 95–105.
  139. Brantley, K.E.; Savin, M.C.; Brye, K.R.; Longer, D.E. Pine Woodchip Biochar Impact on Soil Nutrient Concentrations and Corn Yield in a Silt Loam in the Mid-Southern U.S. Agriculture 2015, 5, 30–47.
  140. 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.
  141. Steiner, C.; Glaser, B.; Geraldes Teixeira, W.; Lehmann, J.; Blum, W.E.H.; Zech, W. Nitrogen retention and plant uptake on a highly weathered central Amazonian Ferralsol amended with compost and charcoal. J. Plant Nutr. Soil Sci. 2008, 171, 893–899.
  142. Nicosia, A.; Pampalone, V.; Ferro, V. Effects of Biochar Addition on Rill Flow Resistance. Water 2021, 13, 3036.
  143. Gwenzi, W.; Chaukura, N.; Noubactep, C.; Mukome, F.N. Biochar-based water treatment systems as a potential low-cost and sustainable technology for clean water provision. J. Environ. Manag. 2017, 197, 732–749.
  144. Rao, M.A.; Simeone, G.D.R.; Scelza, R.; Conte, P. Biochar based remediation of water and soil contaminated by phenanthrene and pentachlorophenol. Chemosphere 2017, 186, 193–201.
  145. Veiga, P.A.D.S.; Schultz, J.; Matos, T.T.D.S.; Fornari, M.R.; Costa, T.G.; Meurer, L.; Mangrich, A.S. Production of high-performance biochar using a simple and low-cost method: Optimization of pyrolysis parameters and evaluation for water treatment. J. Anal. Appl. Pyrolysis 2020, 148, 104823.
  146. Gwenzi, W.; Chaukura, N.; Wenga, T.; Mtisi, M. Biochars as media for air pollution control systems: Contaminant removal, applications and future research directions. Sci. Total. Environ. 2020, 753, 142249.
  147. Blanco-Canqui, H. Biochar and Soil Physical Properties. Soil Sci. Soc. Am. J. 2017, 81, 687–711.
  148. Omondi, M.O.; Xia, X.; Nahayo, A.; Liu, X.; Korai, P.K.; Pan, G. Quantification of biochar effects on soil hydrological properties using meta-analysis of literature data. Geoderma 2016, 274, 28–34.
  149. Ouyang, L.; Wang, F.; Tang, J.; Yu, L.; Zhang, R. Effects of biochar amendment on soil aggregates and hydraulic properties. J. Soil Sci. Plant Nutr. 2013, 13, 991–1002.
  150. Crane-Droesch, A.; Abiven, S.; Jeffery, S.; Torn, M.S. Heterogeneous global crop yield response to biochar: A meta-regression analysis. Environ. Res. Lett. 2013, 8, 044049.
  151. Liu, X.; Zhang, A.; Ji, C.; Joseph, S.; Bian, R.; Li, L.; Pan, G.; Paz-Ferreiro, J. Biochar’s effect on crop productivity and the dependence on experimental conditions—A meta-analysis of literature data. Plant Soil 2013, 373, 583–594.
  152. Yamato, M.; Okimori, Y.; Wibowo, I.F.; Anshori, S.; Ogawa, M. Effects of the application of charred bark of Acacia mangium on the yield of maize, cowpea and peanut, and soil chemical properties in South Sumatra, Indonesia. Soil Sci. Plant Nutr. 2006, 52, 489–495.
  153. Uzoma, K.C.; Inoue, M.; Andry, H.; Fujimaki, H.; Zahoor, A.; Nishihara, E. Effect of cow manure biochar on maize productivity under sandy soil condition. Soil Use Manag. 2011, 27, 205–212.
  154. Alvar-Beltrán, J.; Dao, A.; Marta, A.D.; Heureux, A.; Sanou, J.; Orlandini, S. Farmers’ perceptions of climate change and agricultural adaptation in Burkina Faso. Atmosphere 2020, 11, 0827.
  155. Schmidt, H.P.; Pandit, B.H.; Martinsen, V.; Cornelissen, G.; Conte, P.; Kammann, C.I. Fourfold Increase in Pumpkin Yield in Response to Low-Dosage Root Zone Application of Urine-Enhanced Biochar to a Fertile Tropical Soil. Agriculture 2015, 5, 723–741.
  156. 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.
  157. Alotaibi, K.D. Use of biochar for alleviating negative impact of salinity stress in corn grown in arid soil. Can. J. Soil Sci. 2022, 102, 187–196.
  158. Hung, C.-Y.; Hussain, N.; Husk, B.R.; Whalen, J.K. Ammonia volatilization from manure mixed with biochar. Can. J. Soil Sci. 2022, 102, 177–186.
  159. Shang, X.; Hung, C.-Y.; Husk, B.; Orsat, V.; Whalen, J.K. Wood-based biochar for small fruit production in southern Quebec, Canada. Can. J. Soil Sci. 2022, 102, 89–96.
  160. Glaser, B.; Wiedner, K.; Seelig, S.; Schmidt, H.-P.; Gerber, H. Biochar organic fertilizers from natural resources as substitute for mineral fertilizers. Agron. Sustain. Dev. 2014, 35, 667–678.
  161. Jeffery, S.; Abalos, D.; Prodana, M.; Bastos, A.C.; Van Groenigen, J.W.; Hungate, B.A.; Verheijen, F. Biochar boosts tropical but not temperate crop yields. Environ. Res. Lett. 2017, 12, 053001.
  162. Razzaghi, F.; Obour, P.B.; Arthur, E. Does biochar improve soil water retention? A systematic review and meta-analysis. Geoderma 2019, 361, 114055.
  163. Singh, H.; Northup, B.K.; Rice, C.W.; Prasad, P.V.V. Biochar applications influence soil physical and chemical properties, microbial diversity, and crop productivity: A meta-analysis. Biochar 2022, 4, 8.
  164. Sizmur, T.; Fresno, T.; Akgül, G.; Frost, H.; Moreno-Jiménez, E. Biochar modification to enhance sorption of inorganics from water. Bioresour. Technol. 2017, 246, 34–47.
  165. Curaqueo, G.; Meier, S.; Khan, N.; Cea, M.; Navia, R. Use of biochar on two volcanic soils: Effects on soil properties and barley yield. J. Soil Sci. Plant Nutr. 2014, 14, 911–924.
  166. Agegnehu, G.; Srivastava, A.; Bird, M.I. The role of biochar and biochar-compost in improving soil quality and crop performance: A review. Appl. Soil Ecol. 2017, 119, 156–170.
  167. Sahota, S.; Vijay, V.K.; Subbarao, P.; Chandra, R.; Ghosh, P.; Shah, G.; Kapoor, R.; Vijay, V.; Koutu, V.; Thakur, I.S. Characterization of leaf waste based biochar for cost effective hydrogen sulphide removal from biogas. Bioresour. Technol. 2018, 250, 635–641.
  168. Windeatt, J.H.; Ross, A.B.; Williams, P.T.; Forster, P.M.; Nahil, M.A.; Singh, S. Characteristics of biochars from crop residues: Potential for carbon sequestration and soil amendment. J. Environ. Manag. 2014, 146, 189–197.
  169. O’Laughlin, J.; McElligott, K. Biochar for Environmental Management: Science and Technology, Johannes Lehmann Stephen M. Joseph Earthscan, London UK (2009), 448 p. For. Policy Econ. 2009, 11, 535–536.
  170. Vaccari, F.P.; Baronti, S.; Lugato, E.; Genesio, L.; Castaldi, S.; Fornasier, F.; Miglietta, F. Biochar as a strategy to sequester carbon and increase yield in durum wheat. Eur. J. Agron. 2011, 34, 231–238.
  171. He, Y.; Zhou, X.; Jiang, L.; Li, M.; Du, Z.; Zhou, G.; Shao, J.; Wang, X.; Xu, Z.; Bai, S.H.; et al. Effects of biochar application on soil greenhouse gas fluxes: A meta-analysis. GCB Bioenergy 2016, 9, 743–755.
  172. Hernandez-Soriano, M.C.; Kerré, B.; Kopittke, P.M.; Horemans, B.; Smolders, E. Biochar affects carbon composition and stability in soil: A combined spectroscopy-microscopy study. Sci. Rep. 2016, 6, 25127.
  173. Conte, P.; Bertani, R.; Sgarbossa, P.; Bambina, P.; Schmidt, H.-P.; Raga, R.; Papa, G.L.; Martino, D.F.C.; Meo, P.L. Recent Developments in Understanding Biochar’s Physical–Chemistry. Agronomy 2021, 11, 615.
  174. Trivedi, N.S.; Mandavgane, S.A.; Chaurasia, A. Characterization and valorization of biomass char: A comparison with biomass ash. Environ. Sci. Pollut. Res. 2017, 25, 3458–3467.
  175. Qian, T.; Wang, Y.; Fan, T.; Fang, G.; Zhou, D. A new insight into the immobilization mechanism of Zn on biochar: The role of anions dissolved from ash. Sci. Rep. 2016, 6, 33630.
  176. Xiao, R.; Awasthi, M.K.; Li, R.; Park, J.; Pensky, S.M.; Wang, Q.; Wang, J.J.; Zhang, Z. Recent developments in biochar utilization as an additive in organic solid waste composting: A review. Bioresour. Technol. 2017, 246, 203–213.
  177. Vijay, V.; Shreedhar, S.; Adlak, K.; Payyanad, S.; Sreedharan, V.; Gopi, G.; van der Voort, T.S.; Malarvizhi, P.; Yi, S.; Gebert, J.; et al. Review of Large-Scale Biochar Field-Trials for Soil Amendment and the Observed Influences on Crop Yield Variations. Front. Energy Res. 2021, 9, 710766.
  178. Messiga, A.J.; Hao, X.; Ziadi, N.; Dorais, M. Reducing peat in growing media: Impact on nitrogen content, microbial activity, and CO2 and N2O emissions. Can. J. Soil Sci. 2022, 102, 77–87.
  179. Bieser, J.M.; Al-Zayat, M.; Murtada, J.; Thomas, S.C. Biochar mitigation of allelopathic effects in three invasive plants: Evidence from seed germination trials. Can. J. Soil Sci. 2022, 102, 213–224.
  180. Manirakiza, E.; Ziadi, N.; Luce, M.S.; Hamel, C.; Antoun, H.; Karam, A. Changes in soil pH and nutrient extractability after co-applying biochar and paper mill biosolids. Can. J. Soil Sci. 2022, 102, 27–38.
  181. Ziadi, N.; Zhang, X.; Gagnon, B.; Manirakiza, E. Soil phosphorus fractionation after co-applying biochar and paper mill biosolids. Can. J. Soil Sci. 2022, 102, 53–63.
  182. Hangs, R.; Schoenau, J.; Knight, J. Impact of manure and biochar additions on annual crop growth, nutrient uptake, and fate of 15N-labelled fertilizer in two contrasting temperate prairie soils after four years. Can. J. Soil Sci. 2022, 102, 109–130.
More
Information
Subjects: Soil Science
Contributors MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register : ,
View Times: 146
Revisions: 2 times (View History)
Update Date: 24 Aug 2023
1000/1000