SciELO - Scientific Electronic Library Online

 
vol.38 issue3Relationship between chemical fertilization in sorghum and Melanaphis sacchari/sorghi (Hemiptera: Aphididae) populationsA comparison of two open-source crop simulation models for a potato crop author indexsubject indexarticles search
Home Pagealphabetic serial listing  

Services on Demand

Journal

Article

Indicators

Related links

  • On index processCited by Google
  • Have no similar articlesSimilars in SciELO
  • On index processSimilars in Google

Share


Agronomía Colombiana

Print version ISSN 0120-9965

Agron. colomb. vol.38 no.3 Bogotá Dec. 2020  Epub June 04, 2021

https://doi.org/10.15446/agron.colomb.v38n3.87398 

Soils, fertilization and management of water

Impact of biochar use on agricultural production and climate change. A review

Impacto del uso del biocarbón sobre la producción agrícola y el cambio climático. Una revisión

Sandra Moreno-Riascos1  2 

Thaura Ghneim-Herrera1  * 

1 Departamento de Ciencias Biológicas, Universidad ICESI, Cali (Colombia).

2 Center for Bioinformatics and Photonics-CIBioFi, Universidad del Valle, Cali (Colombia).


ABSTRACT

Biochar is a solid material obtained from the thermal decomposition of biomass of diverse biological origins through a process called pyrolysis. Biochar has great potential for reducing greenhouse gas emissions, sequester carbon in the soil, rehabilitate degraded soils, and reduce dependence on chemical fertilizers in crops. It also improves the physical, chemical, and biological properties of the soil and has a positive effect on plant growth. Given these attributes, there is a growing interest for adopting its use in agriculture, soil and land reclamation, and climate change mitigation. The effects of biochar application can be neutral or positive and will be determined mainly by factors such as the origin of the raw materials, carbonization conditions, frequency of applications, the method of application and dosage. In this review, we offer a detailed examination of the origins of biochar and the technologies used for its production. We examine the various materials that have been used to produce biochars and how they affect their physicochemical characteristics, and we describe their applications in agriculture and climate change mitigation. Finally, we list the guides that describe the standards for the production, characterization, and commercialization of biochar that seek to guarantee the quality of the product and the essential characteristics for its safe use.

Key words: pyrolysis; carbon; biomass; amendment

RESUMEN

El biocarbón es un material sólido obtenido a partir de la descomposición térmica de biomasa de diverso origen biológico mediante un proceso llamado pirólisis. El biocarbón tiene un gran potencial para reducir las emisiones de gases de efecto invernadero, secuestrar el carbono en el suelo, rehabilitar los suelos degradados y reducir la dependencia de los fertilizantes químicos en los cultivos. También mejora las propiedades físicas, químicas y biológicas del suelo y tiene un efecto positivo en el crecimiento de las plantas. Teniendo en cuenta estos atributos, existe un interés creciente en adoptar su uso en la agricultura, la recuperación de suelos y tierras, y en la mitigación del cambio climático. Los efectos de la aplicación del biocarbón pueden ser neutros o positivos y estarán determinados principalmente por factores como el origen de la materia prima, las condiciones de carbonización, la frecuencia de las aplicaciones, el método de aplicación y la dosis. En este artículo ofrecemos un examen detallado de los orígenes del biocarbón y las tecnologías utilizadas para su producción. Examinamos los diversos materiales que se han utilizado para producir biocarbones y cómo éstos afectan a sus características fisicoquímicas, y describimos sus aplicaciones en la agricultura y la mitigación del cambio climático. Por último, enumeramos las guías que describen las normas de producción, caracterización y comercialización del biocarbón, las cuales tratan de garantizar la calidad del producto y las características esenciales para su uso seguro.

Palabras clave: pirólisis; carbono; biomasa; enmienda

Introduction

Biochar is a solid, porous, carbon-enriched compound, produced by high temperature biomass thermal degradation under an inert atmosphere (absence of oxygen) through a thermochemical process known as pyrolysis (Itskos et al., 2016; Basu, 2018; Baskar et al., 2019). This process consists of simultaneous, successive and serial reactions that include dehydration, depolymerization, isomerization, aromatization, decarboxylation, and carbonization (Lee et al., 2019; Xiao et al., 2020). These stages are intertwined (Kan et al., 2016) and cover a complex set of reactions that involve the formation of radical groups (Odinga et al., 2020). Although it is widely considered as an independent biomass conversion technology, pyrolysis actually constitutes the first stage of the gasification and combustion processes (Levin et al., 2016). The main stages of pyrolysis have been generally described as: i) initial evaporation of free moisture, ii) primary decomposition, and iii) secondary reactions (oil cracking and repolymerization) (Uddin et al., 2018; Chen et al., 2019).

The complexity of the pyrolysis process is strongly related to the conditions of manufacturing such as the maximum process temperature, the particle size, the residence time (Zhang et al., 2019), the temperature/heating time relationship, pressure, etc. (Itskos et al., 2016; Leng & Huang, 2018; Szwaja et al., 2018; Sato et al., 2019). During this process, large chains of polymers and inorganic compounds, mainly made up of carbon, hydrogen and oxygen, decompose into smaller molecules in the form of gases, condensable vapors (tars and oils) and solid carbon (Kan et al., 2016). The diversity in composition of the biomass, depending on its origin, added to the conditions described above, will determine the speed and degree of decomposition of each of the components (Tabakaev et al., 2019) as well as their performance and quality (Intani et al., 2016).

Through direct and indirect effects, biochar can improve the physicochemical properties and quality of the soil, increase the productivity of crops, or help mitigate climate change (Cheng et al., 2018; Diatta et al., 2020). That is why biochar has managed to position itself as a key input within the wide range of alternative practices for more sustainable development, thanks to its versatility and potential in different fields such as agriculture or the environment. Some effects derived from the application of biochar to the soil are a reduction of greenhouse gas emissions and the promotion of carbon sequestration in the soil, making possible the rehabilitation of degraded soils (Bis et al., 2018; Gupta et al., 2020). Additionally, biochar reduces dependence on chemical fertilizers in crops by improving the physical, chemical and biological properties of the soil (pH, surface area, cation exchange capacity (CEC), particle density, humidity, and conductivity) (Burrell et al., 2016; Zwart, 2020), reducing the leaching of nutrients, or making a direct or indirect contribution of nutrients that exerts a promoting effect on plant growth and increasing crop yield (Ouyang et al., 2014; Sun et al., 2019). Biochar not only improves the physicochemical properties of the soil, but it can also have a positive impact on its biological properties, such as an increase in the abundance and diversity of beneficial soil microorganisms, since it provides a micro habitat that serves as protection against other pathogenic microorganisms (Liao et al., 2016; Ajema, 2018).

However, it is important to note that the type and magnitude of the effect on soil properties are associated not only with the chemical and physical properties of biochar (Weber & Quicker, 2018), but they can also vary remarkably depending on the type of soil where it is added. Therefore, the characteristics of the biochar and the type of soil are two important factors to consider before its use.

In this review, we provide detailed information about the different technologies, process conditions, and starting materials employed for biochar production and their influence on the final physico-chemical characteristics and practical applications of biochar. We also analyze the impact of biochar use in various applications, in areas such as the mitigation of climate change and agriculture, as well as the international regulations that govern biochar use.

A brief history of biochar

The origin of biochar goes back to the black earths, anthropogenic dark earths (ADE), or Terra Preta de Indio, as they were known by the indigenous populations that inhabited the Brazilian Amazon region (De Oliveira et al., 2020). These lands exhibit completely different properties to most soils in that region (Glaser & Birk, 2012). These soils are widely distributed throughout the region and are characterized by having a high content of nutrients such as carbon (C), phosphorus (P), calcium (Ca), magnesium (Mg), zinc (Zn), manganese (Mn) and organic matter (Barbosa et al., 2020). Its organic component does not come strictly from the accumulation of vegetal cover, but rather it has originated as a result of ancestral human practices as a method for the use of plant and animal bone remains, cooking coal, and waste from fires that were made to replace forests with farming lands (Costa et al., 2017; Macedo et al., 2017). The use of coal for promoting agronomic productivity has been known for centuries (Ogawa & Okimori, 2010), since it retains its stability and properties for hundreds of years. Also, it has proven to be an excellent soil amendment (Ye et al., 2020).

The term biochar was used for the first time in 1999 to describe an activated carbon prepared from sorghum grain and deployed in a process known as reverse-burn gasification (the ChemChar process) (Bapat et al., 1999). Later, it was incorporated into scientific literature on bioenergy to describe charcoal prepared from various crop residues to be used as fuel. The idea of using biochar as an alternative to mitigate climate change was conceived by different authors since 1993, long before the term was born (Paustian et al., 2016). However, it was not until the beginning of the 21st century that Johannes Lehmann associated biochar with the concept of climate change as a way to reduce greenhouse gas emissions (GHG); thus, biochar started to be considered as a soil amendment (Lehmann et al., 2006; Hyland et al., 2010).

Influence of pyrolysis process parameters and type of biomass on biochar characteristics

Biochar can be produced from a variety of biological, raw materials such as crop residues, decaying organic matter of animal origin (livestock manure, pigs, poultry, etc.) (Tag et al., 2016; Pariyar et al., 2020; Tomczyk et al., 2020). Regarding crop residues, the raw biomass is constituted mainly by carbohydrates formed from photosynthesis; the structures and compounds are very diverse depending on the type of plant. These characteristics also influence and significantly determine biochar properties (Li et al., 2020). The yield, ash contents, development and size of the pore, surface area, and even pH are just some of the characteristics that change considerably between the different types of starting material (Tan et al., 2017; Jafri et al., 2018).

Several types of biomass have been used to produce biochar with different purposes (Tab. 1).

TABLE 1 Different raw materials used to produce biochar and some areas of study and potential application. 

Lignocellulose, the principal component of plant biomass, is mainly constituted of three organic polymers: cellulose (40-60% w/w), hemicellulose (20-40% w/w) and lignin (10-25% w/w) (Yang et al., 2007). Each one of these polymers react differently during the heat treatment (Weber & Quicker, 2018), due to their different thermal stabilities that influence not only the required treatment temperature, but also the yield regardless of the chosen conditions (Tag et al., 2016; Uddin et al., 2018).

A large number of studies have shown that there are several interactions between the main components of biomass (Yu et al., 2017); therefore, it is not possible to predict the pyrolysis results simply based on the individual thermal behavior of the three components (Yaashikaa et al., 2019). The interaction between hemicellulose and lignin promotes the production of phenols, while it hinders the generation of hydrocarbons (Cao et al., 2018). Lignin also interacts significantly with cellulose by limiting the polymerization of levoglucosan which reduces the formation of biochar, while the interaction between cellulose and hemicellulose has less effect on the formation and distribution of pyrolysis products (Kan et al., 2016).

Low heating rates and long residence times are usually used to produce biochar. At higher temperatures, the carbon content in the final material increases. However, for vegetal residues high temperatures are not ideal due to the increase in ash contents (Rafiq et al., 2016). As temperature increases, water evaporation rises. Volatile compounds are then released, which, in turn, enlarges the relative fixed carbon ratio of the solid compared to the raw biomass (Jouhara et al., 2018). At the same time, there are variations in the carbon, hydrogen, and oxygen contents that lead to some physicochemical changes such as polarity reduction, aromaticity increase, and biochar hydrophobicity (Lehmann et al, 2009; Gray et al, 2014) 500°C, and 620°C. The surface area as well as pore volume also increase slightly due to the release of volatiles during carbonization. This process reconfigures the structure of lignocellulose compounds, giving rise to the formation of structures shaped as canals (Sigmund et al., 2017).

The thermal decomposition of the organic components in the biomass begins between 350-550°C with the formation of two structural fractions composed by the stacking of crystalline sheets of poly-aromatic graphene that grow laterally, and by an amorphous fraction of randomly organized aromatic structures (Weber & Quicker, 2018). Both of these fractions are associated with ring shaped carbon bonds of benzene types with oxygen or hydrogen (Kwiatkowski & Kalderis, 2020), that give rise to layers, forming the structure of a lattice or slit that originates multiple spaces corresponding to the pores. These bonds between aromatic structures of C-O and C-H determine the stability of biochar and are used as a parameter to measure the degree of aromaticity (Choudhary et al., 2019). They also contain oxidized and aliphatic carbon structures that are easily degradable.

The dominant process is carbonization, that occurs at between the 600-800°C, and causes the removal of most non-carbonaceous atoms such as calcium (Ca), magnesium (Mg), and potassium (K) (Zhang et al., 2017; Chatterjee et al., 2020). Hydrogen (H), oxygen (O), phosphorus (P), and sulfur (S) are mainly located within aromatic rings as het-eroatoms. The presence of heteroatoms contributes greatly to the reactivity of biochar and the chemical heterogeneity of its surface (Cheng & Li, 2018).

As a result, biochar with low H/C ratios, that corresponds to a higher degree of carbonization, contain fewer functional groups and more aromatic structures than biochar produced at low temperatures. O/C ratios are much higher than those present in mineral coal that also originates from biomass, but that is formed by geological processes in periods that include geological scales (Bakshi et al., 2020).

The pH of biochar also depends on the production parameters and type of biomass, which results in products with a wide range of pH values (Tag et al., 2016). When biochar is produced at temperatures below 400°C, it reaches a pH below 7, but at temperatures above 500°C, the pH can vary from 7 to more alkaline (Tomczyk et al., 2020).

Biochar production technologies

Within the diverse thermochemical processes that transform biomass into biochar, four methods are used: slow pyrolysis, fast pyrolysis, flash pyrolysis, and gasification (Novak et al., 2010; Ippolito et al, 2012).

Slow pyrolysis, or carbonization, is characterized by long residence times (hours to days) at relatively low temperatures (~300-700°C) and materials with different particle sizes (5-50 mm) (Claoston et al., 2014). In this type of pyrolysis, the thermal decomposition of biomass occurs at a very low heating rate (temperature/time), using enough time to maximize the solid yields thanks to repolymerization reactions. The previous method is used mainly to produce the solid fraction or coal, although it can also be gasified to obtain hydrogen-rich gas.

Fast pyrolysis generally involves high heating rates (>10-200°C sec-1) and short residence times (0.5-10 sec) (Claoston et al., 2014). It is used to obtain a high yield of liquid products, such as bio-oil, in relation to the dry biomass base, which can be as high as 50-70% by weight.

Finally, the pyrolysis process known as flash is characterized by very high heating rates (103-104°C sec-1) and much shorter residence times (<0.5 sec), resulting in very high bio-oil yields that can reach up to 75-80% by weight (Gilbe et al., 2008; Gao et al., 2011; Enders et al., 2012). In comparison to residual biomass, these materials offer greater chemical and biological decomposition resistance (Zimmerman, 2010) as well as a greater number of aromatic structures (Baldock & Smernik, 2002).

Biochar quality standards

In order to guarantee the safety of the use of biochar as a soil amendment to the market and especially consumers, the members of the International Biochar Initiative (IBI) in conjunction with research centers, scientists, farmers, and producers worldwide, developed guidelines and policies that provide the rules for the characterization of biochar products. All the standards are based on a series of guiding principles and the follow-up of guidelines consistent with the best practices for the development of standards such as the International Standards Organization (ISO), ASTM International, and the Institute of Electrical and Electronics Engineers (IEEE). Compliance with these standards by manufacturers is not mandatory. However, the IBI has implemented a certification program to expand the biochar industry since 2013. Through this program the producer can certify that the product complies with quality standards and that it is safe for use in soils. In addition, the IBI Certified TM quality seal is awarded, which is recognized worldwide (International Biochar Initiative, 2015).

At the same time, Europe has the European Biochar Certificate (EBC) that applies a system for the sustainable production of biochar based on the latest and most innovative research and practices. These practices aim to guarantee the production control and quality of biochar with processes that have scientific and legal support and that are economically viable and practically applicable. Unlike the IBI standards, the European certificate is mandatory for all producers (European Biochar Foundation, 2016). Currently, there are two documents approved by the members of the IBI and EBC that contain the technical programs, policies, and guidelines: IBI Biochar standards (version 2.0) and European biochar certificate guidelines (version 6.4E).

Biochar in agriculture

Great interest has been generated on biochar and its application to soils due to the similarity between biochar particles and those found in Terra Preta de Indio (Lehmann et al., 2011). Several studies have shown that the physical and chemical properties of biochar favor adsorption thanks to its high specific surface area compared to other organic amendments, thus improving the availability of nutrients in the soil (Bonanomi et al., 2017; Yadav et al, 2019).

The pH is one of the most important soil parameters because of its great influence on biological, chemical, physical, and geological processes, and because it is related to soil fertility (Neina, 2019). Changes in pH trigger a series of modifications in the soil environment, affecting the availability of nutrients for plant growth and microbial activity, or accelerating biogeochemical processes in the rhizosphere (Ducey et al., 2015). The addition of biochar as a soil amendment has become a common practice, as it improves soil quality by reducing soil acidity. Due to its porous matrix structure, it can also improve soil structure by forming micro aggregates, increase water retention, improve solubilization, nutrient retention and transport, enhance the soil quality index and reduce nutrient leaching (Martinsen et al., 2014; Wang et al., 2014; Pratiwi et al., 2016; Jeffery et al, 2017; Oladele, 2019).

Some properties of biochar, such as high surface charge density, large surface area, internal porosity, and the presence of polar and non-polar surface sites play an important role in the liming effect (Shetty & Prakash, 2020). The effect of pH on acidic soils is probably due to their alkalinity and high buffering capacity (Juriga & Simansky, 2019). The basic cations (e.g., K, Ca, Mg and Si) that are contained in biochar as carbonates or oxides can reduce the acidity by increasing the reaction of the exchangeable basic cations. This process is carried out through the functional groups on the surface of biochar as COO- and O- with the H+ or Al+3 ions in the soil, which would cause an increase in the soil pH (Hansen et al., 2016; Dai et al., 2017; Zong et al, 2018).

Another physicochemical property of the soil associated with its fertility is cation exchange capacity (CEC), which refers to the capacity of soils to retain cations interchangeably through adsorption (Duraes et al., 2018; Zhang et al., 2018). The CEC of biochar is mainly a consequence of the temperature conditions at which it is produced (Tan et al., 2017; Leng & Huang, 2018). The carboxylic groups formed in the bridges of the aromatic nuclei of biochar are responsible for the increase of the CEC and reactivity (Zhang et al., 2018). After biochar is added to the soil, the charges of its functional groups are generally positive, but over time the functional groups on the surface oxidize and generate more negative charges relative to the positive ones, increasing the CEC. Adsorption of highly oxidized organic matter on biochar surfaces is another process that affects CEC (Tomczyk et al., 2020). When biochar is exposed to oxygen and water, more functional groups can be generated on the surface through oxidation, thus increased CEC is attained (Hue, 2020). However, CEC is highly variable depending on the surface chemistry of biochar and tends to change once it is incorporated due to interactions with the environment (Hailegnaw et al, 2019).

These processes can directly or indirectly influence the improvement of other physical and chemical soil properties (Karimi et al., 2020). A positive effect of biochar on the productivity of the crops has been shown (Jin et al., 2016; He et al., 2017; Masud et al., 2020). The most significant changes have been observed in acid soil (Diatta et al., 2020), which suggests that its yield effect is similar to that produced by liming, accompanied by better water retention (Ahmed et al., 2016). The water retention capacity is mainly due to two factors: the large internal surface and the high number of residual pores in the biochar, where water is retained by capillarity, improving soil aggregation and structure. This increases the general porosity of the soil and the water content, leading to a decrease in the mobility of water, reducing water stress in plants (Batista et al, 2018).

Biochar improves the structure, porosity and aggregation of the soil, facilitating tillage (Du et al., 2018) and favoring the availability of nutrients (Karimi et al., 2020). Nutrient composition and availability from biochar depend on biochar raw material and pyrolysis conditions (Purakayastha et al., 2019). In addition, various factors affect the nutrient status in biochar-treated soils including the soil and feedstock type, pyrolysis temperature, and addition rate (Yu et al., 2019). Biochar can directly provide higher nutrient content such as P, K, Ca, Mg (Gao et al., 2017). Also, it can influence soil microbial activity due to changes in labile carbon and soil properties. The activity of soil microbes can significantly affect soil organic matter decomposition and nutrient cycling (Yu et al., 2019). The soil microbial biomass and enzyme activities have major impacts on the soil nutrient status and crop productivity. Another promoted benefit of biochar application to soil is reduced nutrient leaching (Zhao et al., 2019; Karimi et al., 2020).

Therefore, biochar does not act by means of a single effect. It influences different, interconnected soil physicochemical and biological properties. A highly variable response has been shown in the soil biota, reporting increases, decreases (Kolton et al., 2011; Li et al., 2018), or no significant effects (Lehmann et al., 2011). These interactions between microorganisms and biochar are also controlled by multiple environmental factors such as types and rates of biochar amendment, soil type, land use, and vegetation types (Gorovtsov et al., 2020).

Biochar as a matrix carrier for plant growth-promoting microorganisms

Peat is the most widely used inoculum carrier in the world; however, its availability is limited (Sahu & Brahmaprakash, 2016). Exploration of other matrix alternatives generated by more sustainable processes and that compete biologically and economically with the materials currently used could be beneficial to the biofertilizer industry (Herrmann & Lesueur, 2013; Flores-Félix et al., 2019; Saeid & Chojnacka, 2019). Biochar has the potential to be used as an alternative vehicle in the development of new biofertilizers. Its use as a carrier of plant growth-promoting microorganisms could favor both the proliferation of the microorganism of interest immobilized in it and the abundance of indigenous microorganisms (Ajema, 2018). However, it is important to consider that the raw material, the pyrolysis temperature, the degree of oxidation, and the size of the pores obtained during the production of biochar affect the characteristics of the final product and, therefore, its effect on microorganisms. The effects of inoculated biochar also vary according to the environmental conditions that directly affect the viability of microorganisms (Palansooriya et al, 2019).

The effects of biochar incorporation on soil microorganisms can be direct through the contribution of labile organic matter that provides a source of energy and nutrients. They can also be indirect, through changes in some soil properties, such as pH or porosity (Gorovtsov et al., 2020). The porous structure of biochar can act as a habitat for microorganisms. It can also increase the soil surface, which is able to retain water and nutrients favoring the growth of soil biota and rhizosphere microbiome (Jenkins et al., 2017; Husna et al., 2019).

The addition of biochar can increase mineralization through its contribution of labile organic matter. In general, biochar obtained through slow pyrolysis contains a greater amount of labile organic matter and tends to cause a greater increase in soil biota than more recalcitrant biochar (Cross & Sohi, 2011; Hardy et al, 2019).

Hale et al. (2014) show that survival but not abundance of Enterobacter cloacae (UW5), a well-studied bacterium that produces indole acetic acid (IAA) in a tryptophan-dependent manner, increases in soils amended with any of five different types of biocarbon (palm leaves, pine wood, coconut shell, pistachio shell, and fruit pits) when compared to peat and vermiculite. Ghazi (2017) evaluates the effect of Rhizobium-based formulations developed with biochar from rice husks on bean seeds, which were compared with peat and vermiculite materials. The biochar-based treatment records a maximum carrier survival population after 180 d of inoculation. In addition, a slight decrease in pH was recorded at the end of the storage period.

Biochar derived from acacia wood and inoculated with Azospirillum lipoferum (AZ 204) was tested at different dose levels (0, 5, 10, 15, 20, 25 t ha-1) and its effect on maize growth was evaluated. A significant increase was observed in Azospirillum and other diazotroph populations in the rhizosphere after the application of biochar at all the stages of crop growth. Furthermore, a significant increase of native mycorrhizal colonization was observed in response to biochar-Azospirillum application (Saranya et al., 2011).

Additionally, combinations of bacteria have been evaluated in the formulation of biofertilizers based on biochar, with the aim of finding sets of microorganisms that enhance plant growth through an interactive and harmonious work. The microbial consortium formed by endophytic bacteria (Brevibacillus, Enterobacter, Kytococcus, Pantoea, Pseudomonas, Serratia, and Stenotrophomonas) and fungi (Cutaneotrichosporon, Mucor, and Wickerhamomyces) isolated from rapeseed (Brassica napus) and barley (Hor-deum vulgare) were inoculated into a biochar derived from processed wood waste to evaluate its potential as a carrier of plant growth promoting microorganisms (PGPM) and its growth effect on barley seedlings. A synergistic effect of the microbial consortium with the biochar on root growth was observed. A decrease in urease activity in the soil was also found, without causing a negative effect on seedling development during the study (Vecstaudza et al, 2017).

Biochar has been recognized as a carrier matrix not only for microorganisms with a plant growth-promoting effect, but also as an alternative for the immobilization of microorganisms used in bioremediation. For example, biochar was used as a carrier of bacterial consortia in soils contaminated with cypermethrin, an insecticide for agricultural use (Liu et al, 2017).

Biochar can also play an important role as a carrier of bacteria used in biological control. In sugarcane, this strategy showed positive results for the control of the pathogen Macrophomina phaseolina by the antagonistic rhizobacterium Paenibacillus illinoisensis RH-3 (Shahjahan et al., 2018) . That study demonstrated that biochar derived from sugarcane bagasse is a better carrier of P. illinoisensis RH-3, compared to peat and lignite. The neutral pH of biochar together with its porosity and high availability of nutrients seemed to be determinants of its positive effect (Shahjahan et al., 2018).

All research on the potential of biochar as a carrier of microbial inoculum supports its use as an alternative to classical chemical fertilization. However, further studies are needed to elucidate the synergistic effects of biochar and soil microorganisms on plant growth promotion, their mechanisms of action, and the interactions between microorganisms and their microhabitats. Finally, such effects must also be validated under field conditions.

Biochar in climate change mitigation

Biochar was first recognized for its potential as a tool to counteract climate change in 2009 (Lehmann et al., 2009). Its porous structure and high internal surface area give it the capacity to absorb soluble organic compounds, gases, some inorganic compounds, and different pollutants thanks to the diversity of functional groups on its surface (Mendes et al., 2018). Biochar has been evaluated for the removal of heavy metals in contaminated waters such as lead (Pb), copper (Cu), zinc (Zn) and cadmium (Cd) (Shim et al., 2015; Wang et al., 2019; Zhao et al., 2019). The type of biomass and the temperature of the process are very influential for the absorption capacity of biochar. For example, materials produced at high temperatures result in materials with a larger surface area, higher carbon content and microporosity, leading to increased efficiency in removing compounds such as trichloroethylene (TCE), benzene (C6H6) and nitrobenzene (C6H5NO2) (Zhou et al., 2010; Ahmad et al, 2012).

The application of biochar to soils has been proposed as a strategy to reduce the concentration of CO2 in the atmosphere, as it would serve as a long-term carbon sink. Its recalcitrant nature, i.e., its resistance to degradation due to all the carbon present in its structure, gives it the ability to remain in soils for hundreds or thousands of years (Woolf et al., 2010) and to reduce the rate at which carbon fixed by photosynthesis returns to the atmosphere. It has also been suggested that carbon retention can be much greater when soil carbon stocks increase (Ding et al., 2018; Ventura et al., 2019) . However, some studies have found that many of the practices used to sequester carbon in the soil are often offset by increased greenhouse gas emissions (Shen et al., 2014; Zhang et al., 2019) and that soils show a low potential for carbon accumulation (Smith, 2016; Sharma, 2018; Gupta et al., 2020). However, among the possible strategies to remove CO2 from the atmosphere, biochar stands out in this area and currently represents the most promising alternative.

Biochar application can also reduce the emissions of nitrous oxide (N2O) and sometimes, the flow CO2 in soil, generating an additional tool to mitigate their effects (Woolf et al., 2010). However, the results have not always been positive in terms of CO2 reductions (Spokas et al., 2010).

Regarding the potential of biochar to reduce nitrous oxide (N2O) emissions, several hypotheses have been put forward on how biochar could interact with denitrification, by directly stimulating or suppressing the amount of N that passes into the gaseous form. These hypotheses relate this effect to several different biochar properties, soil type, and environmental conditions such as temperature and precipitation (Dicke et al., 2015; Lan et al., 2019). Although to date the impact of biochar on denitrification has not yielded conclusive results (Cayuela et al., 2014; Ameloot et al., 2016; Kammann et al, 2017), it appears that the magnitude and mechanisms associated with biochar inhibition of N2O depend on the dose and application frequency (Liu et al., 2020).

Biochar-induced changes in the composition or activity of the microbial community not only affect nutrient cycles and plant growth but also the soil organic matter cycle (Ouyang et al., 2014; Hardy et al., 2019; Gorovtsov et al., 2020). While research under controlled conditions has shown that biochar can affect N2O emissions from the soil (Taghizadeh-Toosi et al., 2011; Schimmelpfennig et al., 2014; Kammann et al., 2017), laboratory results cannot be extrapolated to what is expected in the field. In field trials, often no statistical differences are observed between biochar and control treatments. Despite the extensive literature published on the subject in recent years, it is still very unpredictable whether a type of biochar will be effective in mitigating N2O emissions in a field with crops. Factors such as application rate, biochar characteristics, soil type, environmental conditions, less homogeneous particle distribution, and greater soil (and plant) heterogeneity in the fields result in a high variability in N2O flows (Huppi et al., 2015). Therefore, most research efforts are now directed towards achieving the greatest reductions in N2O emissions by selecting the most efficient biochars and analyzing the mechanisms involved.

Conclusions

The addition of biochar to the soil provides different benefits derived from changes in the physical, chemical, and biological characteristics of the soil, which in turn, promote plant growth and productivity by increasing the availability of nutrients and pH. The porosity of biochar, along with water and nutrient retention, can promote the establishment and increase the abundance and diversity of microbiota which, in turn, promote plant growth.

The potential of biochar for carbon sequestration and its ability to reduce greenhouse gas emissions make it a very attractive alternative to counteract the adverse effects of climate change. However, further research is needed in this field to elucidate its long-term effects.

The wide availability of biochar as well as the varied responses in crops that will depend not only on the crop species but also on the type of soil, make it difficult to establish standard parameters regarding the dose and form of biochar application. For this reason, it is necessary to expand research in this field, since the wide availability of raw materials and the different technologies available for pyrolysis offer a wide range of opportunities for the production, marketing, and profitable use of biochar in Colombian agriculture.

Acknowledgments

SMR MSc studies were financially supported by a grant from Universidad Icesi and the CIBioFi Project, funded by the Colombian Science, Technology and Innovation Fund - General Royalties System (Fondo CTeI- SGR), Gobernación del Valle del Cauca, and Ministerio de Ciencia, Tecnologia e Innovación (Minciencias), under contract No. BPIN 2013000100007. We also kindly thank Franklin Gerardo Ojeda for his revision of the manuscript.

Literature cited

Ahmad, M., Lee, S. S., Dou, X., Mohan, D., Sung, J. K., Yang, J. E., & Ok, Y. S. (2012). Effects of pyrolysis temperature on soybean stover- and peanut shell-derived biochar properties and TCE adsorption in water. Bioresource Technology, 118, 536-544. https://doi.org/10.1016/j.biortech.2012.05.042Links ]

Ahmed, A., Kurian, J., & Raghavan, V. (2016). Biochar influences on agricultural soils, crop production, and the environment: a review. Environmental Reviews, 24(4), 495-502. https://doi.org/10.1139/er-2016-0008Links ]

Ajema, L. (2018). Effects of biochar application on beneficial soil organism review. International Journal of Research Studies in Science, Engineering and Technology, 5(5), 9-18. https://doi.org/10.13140/RG.2.2.15186.66247Links ]

Ameloot, N., Maenhout, P., De Neve, S., & Sleutel, S. (2016). Biochar-induced N2O emission reductions after field incorporation in a loam soil. Geoderma, 267, 10-16. https://doi.org/10.1016/j.geoderma.2015.12.016Links ]

Bakshi, S., Banik, C., & Laird, D. A. (2020). Estimating the organic oxygen content of biochar. Scientific Reports, 10(1), Article 13082. https://doi.org/10.1038/s41598-020-69798-yLinks ]

Baldock, J. A., & Smernik, R. J. (2002). Chemical composition and bioavailability of thermally altered Pinus resinosa (Red pine) wood. Organic Geochemistry, 33(9), 1093-1109. https://doi.org/10.1016/S0146-6380(02)00062-1Links ]

Bapat, H., Manahan, S. E., & Larsen, D. W. (1999). An activated carbon product prepared from milo (Sorghum vulgare) grain for use in hazardous waste gasification by ChemChar cocurrent flow gasification. Chemosphere, 39(1), 23-32. https://doi.org/10.1016/S0045-6535(98)00585-2Links ]

Barbosa, J. Z., Motta, A. C. V., Corrêa, R. S., Melo, V. de F., Mu-niz, A. W., Martins, G. C., Silva, L. de C. R., Teixeira, W. G., Young, S. D., & Broadley, M. R. (2020). Elemental signatures of an Amazonian Dark Earth as result of its formation process. Geoderma, 361, Article 114085. https://doi.org/10.1016/j.geoderma.2019.114085Links ]

Baskar, G., Kalavathy, G., Aiswarya, R., & Abarnaebenezer Selvakumari, I. (2019). Advances in bio-oil extraction from nonedible oil seeds and algal biomass. In K. Azad (Ed.), Advances in eco-fuels for a sustainable environment, a volume in Woodhead Publishing series in energy (pp. 187-210). Woodhead Publishing. https://doi.org/10.1016/B978-0-08-102728-8.00007-3Links ]

Basu, P. (2018). Biomass gasification, pyrolysis and torrefaction, practical design and theory (3rd ed.). Academic Press. https://doi.org/10.1016/B978-0-12-812992-0.00005-4Links ]

Batista, E. M. C. C., Shultz, J., Matos, T. T. S., Fornari, M. R., Ferreira, T. M., Szpoganicz, B., de Freitas, R. A., & Mangrich, A. S. (2018). Effect of surface and porosity of biochar on water holding capacity aiming indirectly at preservation of the Amazon biome. Scientific Reports, 8(1), Article 10677. https://doi.org/10.1038/s41598-018-28794-zLinks ]

Bis, Z., Kobylecki, R., Scislowska, M., & Zarzycki, R. (2018). Biochar - potential tool to combat climate change and drought. Ecohydrology & Hydrobiology, 18(4), 441-453. https://doi.org/10.1016/j.ecohyd.2018.11.005Links ]

Bonanomi, G., Ippolito, F., Cesarano, G., Nanni, B., Lombardi, N., Rita, A., Saracino, A., & Scala, F. (2017). Biochar as plant growth promoter: better off alone or mixed with organic amendments? Frontiers in Plant Science, 8, Article 1570. https://doi.org/10.3389/fpls.2017.01570Links ]

Burrell, L. D., Zehetner, F., Rampazzo, N., Wimmer, B., & Soja, G. (2016). Long-term effects of biochar on soil physical properties. Geoderma, 282, 96-102. https://doi.org/10.1016/j.geoderma.2016.07.019Links ]

Cao, W., Li, J., Martí-Rosselló, T., & Zhang, X. (2018). Experimental study on the ignition characteristics of cellulose, hemicellulose, lignin and their mixtures. Journal of the Energy Institute, 92(5), 1303-1312. https://doi.Org/10.1016/j.joei.2018.10.004Links ]

Cayuela, M. L., van Zwieten, L., Singh, B. P., Jeffery, S., Roig, A., & Sánchez-Monedero, M. A. (2014). Biochar's role in mitigating soil nitrous oxide emissions: a review and meta-analysis. Agriculture, Ecosystems and Environment, 191, 5-16. https://doi.org/10.1016/j.agee.2013.10.009Links ]

Chatterjee, R., Sajjadi, B., Chen, W. Y., Mattern, D. L., Hammer, N., Raman, V., & Dorris, A. (2020). Effect ofpyrolysis temperature on physicochemical properties and acoustic-based amination of biochar for efficient CO2 adsorption. Frontiers in Energy Research, 8, Article 85. https://doi.org/10.3389/fenrg.2020.00085Links ]

Chen, C., Liu, G., An, Q., Lin, L., Shang, Y., & Wan, C. (2020). From wasted sludge to valuable biochar by low temperature hydrothermal carbonization treatment: insight into the surface characteristics. Journal of Cleaner Production, 263, Article 121600. https://doi.org/10.1016/j.jclepro.2020.121600Links ]

Chen, W. H., Wang, C. W., Ong, H. C., Show, P. L., & Hsieh, T. H. (2019). Torrefaction, pyrolysis and two-stage thermodegradation of hemicellulose, cellulose and lignin. Fuel, 258, Article 116168. https://doi.org/10.1016/j.fuel.2019.116168Links ]

Cheng, F., & Li, X. (2018). Preparation and application of biochar-based catalysts for biofuel production. Catalysts, 8(9), Article 346. https://doi.org/10.3390/catal8090346Links ]

Cheng, N., Peng, Y., Kong, Y., Li, J., & Sun, C. (2018). Combined effects of biochar addition and nitrogen fertilizer reduction on the rhizosphere metabolomics of maize (Zea mays L.) seedlings. Plant and Soil, 433, 19-35. https://doi.org/10.1007/s11104-018-3811-6Links ]

Cheng, Q., Huang, Q., Khan, S., Liu, Y., Liao, Z., Li, G., & Ok, Y. S. (2016). Adsorption of Cd by peanut husks and peanut husk biochar from aqueous solutions. Ecological Engineering, 87, 240-245. https://doi.org/10.1016/j.ecoleng.2015.11.045Links ]

Choudhary, T. K., Khan, K. S., Hussain, Q., Ahmad, M., & Ashfaq, M. (2019). Feedstock-induced changes in composition and stability of biochar derived from different agricultural wastes. Arabian Journal of Geosciences, 12, Article 617. https://doi.org/10.1007/s12517-019-4735-zLinks ]

Claoston, N., Samsuri, A. W., Ahmad Husni, M. H., & Mohd Amran, M. S. (2014). Effects of pyrolysis temperature on the physicochemical properties of empty fruit bunch and rice husk biochars. Waste Management and Research, 32(4), 331-339. https://doi.org/10.1177/0734242X14525822Links ]

Costa, A. da R., Silva Júnior, M. L., Kern, D. C., Ruivo, M. de L. P., & Marichal, R. (2017). Forms of soil organic phosphorus at black earth sites in the Eastern Amazon. Revista Ciência Agronômica, 48(1), 1-12. https://doi.org/10.5935/1806-6690.20170001Links ]

Cross, A., & Sohi, S. P. (2011). The priming potential of biochar products in relation to labile carbon contents and soil organic matter status. Soil Biology and Biochemistry, 43(10), 2127-2134. https://doi.org/10.1016/j.soilbio.2011.06.016Links ]

Dai, Z., Zhang, X., Tang, C., Muhammad, N., Wu, J., Brookes, P. C., & Xu, J. (2017). Potential role of biochars in decreasing soil acidification - A critical review. Science of The Total Environment, 581-582, 601-611. https://doi.org/10.1016/j.scitotenv.2016.12.169Links ]

De Oliveira, E. A., Marimon-Junior, B. H., Marimon, B. S., Iriarte, J., Morandi, P. S., Maezumi, S. Y., Nogueira, D. S., Aragão, L. E. O. C, da Silva, I. B., & Feldpausch, T. R. (2020). Legacy of Amazonian Dark Earth soils on forest structure and species composition. Global Ecology and Biogeography, 29(9), 1458-1473. https://doi.org/10.1111/geb.13116Links ]

Diatta, A. A., Fike, J. H., Battaglia, M. L., Galbraith, J. M., & Baig, M. B. (2020). Effects of biochar on soil fertility and crop productivity in arid regions: a review. Arabian Journal of Geosciences, 13(14), Article 595. https://doi.org/10.1007/s12517-020-05586-2Links ]

Dicke, C., Andert, J., Ammon, C., Kern, J., Meyer-Aurich, A., & Kaupenjohann, M. (2015). Effects of different biochars and digestate on N2O fluxes under field conditions. Science of The Total Environment, 524-525, 310-318. https://doi.org/10.1016/j.scitotenv.2015.04.005Links ]

Ding, F., Van Zwieten, L., Zhang, W., Weng, Z. H., Shi, S., Wang, J., & Meng, J. (2018). A meta-analysis and critical evaluation of influencing factors on soil carbon priming following biochar amendment. Journal of Soils and Sediments, 18, 1507-1517. https://doi.org/10.1007/s11368-017-1899-6Links ]

Du, Z., Xiao, Y., Qi, X., Liu, Y., Fan, X., & Li, Z. (2018). Peanut-shell biochar and biogas slurry improve soil properties in the North China Plain: a four-year field study. Scientific Reports, 8, Article 13724. https://doi.org/10.1038/s41598-018-31942-0Links ]

Ducey, T. F., Novak, J. M., & Johnson, M. G. (2015). Effects of biochar blends on microbial community composition in two coastal plain soils. Agriculture, 5(4), 1060-1075. https://doi.org/10.3390/agriculture5041060Links ]

Durâes, N., Novo, L. A. B., Candeias, C., & da Silva, E. F. (2018). Distribution, transport and fate of pollutants. In A. C. Duarte, A. Cachada, & T. Rocha-Santos (Eds.), Soil pollution, from monitoring to remediation (pp. 29-57). Academic Press. https://doi.org/10.1016/B978-0-12-849873-6.00002-9Links ]

Enders, A., Hanley, K., Whitman, T., Joseph, S., & Lehmann, J. (2012). Characterization of biochars to evaluate recalcitrance and agronomic performance. Bioresource Technology, 114, 644-653. https://doi.org/10.1016/j.biortech.2012.03.022Links ]

European Biochar Foundation. (2016). Guidelines for a sustainable production of biochar. European Biochar Foundation (EBC). https://doi.org/10.13140/RG.2.1.4658.7043Links ]

Flores-Félix, J. D., Menéndez, E., Rivas, R., & Velazquez, M. de la E. (2019). Future perspective in organic farming fertilization: management and product. In S. Chandran, M. R. Unni, & S. Thomas (Eds.), Organic farming global perspectives and methods (pp. 269-315). Woodhead Publishing. https://doi.org/10.1016/B978-0-12-813272-2.00010-0Links ]

Gao, L., Wang, R., Shen, G., Zhang, J., Meng, G., & Zhang, J. (2017). Effects of biochar on nutrients and the microbial community structure of tobacco-planting soils. Journal of Soil Science and Plant Nutrition, 17(4), 884-896. https://doi.org/10.4067/S0718-95162017000400004Links ]

Gao, X., & Wu, H. (2011). Biochar as a fuel: 4. Emission behavior and characteristics of PM1 and PM10 from the combustion of pulverized biochar in a drop-tube furnace. Energy & Fuels, 25(6), 2702-2710. https://doi.org/10.1021/ef200296uLinks ]

Ghazi, A. A. (2017). Potential for biochar as an alternate carrier to peat moss for the preparation of Rhizobia bio inoculum. Microbiology Research Journal International, 18(4), 1-9. https://doi.org/10.9734/MRJI/2017/30828Links ]

Gilbe, C., Öhman, M., Lindström, E., Boström, D., Backman, R., Samuelsson, R., & Burvall, J. (2008). Slagging characteristics during residential combustion of biomass pellets. Energy & Fuels, 22(5), 3536-3543. https://doi.org/10.1021/ef800087xLinks ]

Glaser, B., & Birk, J. J. (2012). State of the scientific knowledge on properties and genesis of Anthropogenic Dark Earths in Central Amazonia (terra preta de índio). Geochimica et Cosmochimica Acta, 82, 39-51. https://doi.org/10.1016/j.gca.2010.11.029Links ]

Gorovtsov, A. V., Minkina, T. M., Mandzhieva, S. S., Perelomov, L. V., Soja, G., Zamulina, I. V., Rajput, V. D., Sushkova, S. N., Mohan, D., & Yao, J. (2020). The mechanisms of biochar interactions with microorganisms in soil. Environmental Geochemistry and Health, 42, 2495-2518. https://doi.org/10.1007/s10653-019-00412-5Links ]

Gray, M., Johnson, M. G., Dragila, M. I., & Kleber, M. (2014). Water uptake in biochars: the roles of porosity and hydrophobicity. Biomass and Bioenergy, 61, 196-205. https://doi.org/10.1016/j.biombioe.2013.12.010Links ]

Gupta, D. K., Gupta, C. K., Dubey, R., Fagodiya, R. K., Sharma, G,., A,K., Noor Mohamed, M. B., Dev, R., & Shukla, A. K. (2020). Role of biochar in carbon sequestration and greenhouse gas mitigation. In J. S. Singh & C. Singh (Eds.), Biochar applications in agriculture and environment management (pp. 141-165). Springer International Publishing. https://doi.org/10.1007/978-3-030-40997-5_7Links ]

Hailegnaw, N. S., Mercl, F., Pracke, K., Száková, J., & Tlustos, P. (2019). Mutual relationships of biochar and soil pH, CEC, and exchangeable base cations in a model laboratory experiment. Journal of Soils and Sediments, 19, 2405-2416. https://doi.org/10.1007/s11368-019-02264-zLinks ]

Hale, L., Luth, M., Kenney, R., & Crowley, D. (2014). Evaluation of pinewood biochar as a carrier of bacterial strain Enterobacter cloacae UW5 for soil inoculation. Applied Soil Ecology, 84, 192-199. https://doi.org/10.1016/j.apsoil.2014.08.001Links ]

Hansen, V., Múller-Stõver, D., Munkholm, L. J., Peltre, C., Haug-gaard-Nielsen, H., & Jensen, L. S. (2016). The effect of straw and wood gasification biochar on carbon sequestration, selected soil fertility indicators and functional groups in soil: an incubation study. Geoderma, 269, 99-107. https://doi.org/10.1016/j.geoderma.2016.01.033Links ]

Hardy, B., Sleutel, S., Dufey, J. E., & Cornelis, J. T. (2019). The long-term effect of biochar on soil microbial abundance, activity and community structure Is overwritten by land management. Frontiers in Environmental Science, 7, Article 110. https://doi.org/10.3389/fenvs.2019.00110Links ]

He, Y., Zhou, X., Jiang, L., Li, M., Du, Z., Zhou, G., Shao, J., Wang, X., Xu, Z., Hosseini-Bai, S., Wallace, H., & Xu, C. (2017). Effects of biochar application on soil greenhouse gas fluxes: a meta-analysis. GCB Bioenergy, 9(4), 743-755. https://doi.org/10.1111/gcbb.12376Links ]

Herrmann, L., & Lesueur, D. (2013). Challenges of formulation and quality of biofertilizers for successful inoculation. Applied Microbiology and Biotechnology, 97(20), 8859-8873. https://doi.org/10.1007/s00253-013-5228-8Links ]

Huang, L. Q., Fu, C., Li, T. Z., Yan, B., Wu, Y., Zhang, L., Ping, W., Yang, B. R., & Chen, L. (2020). Advances in research on effects of biochar on soil nitrogen and phosphorus. IOP Conference Series: Earth and Environmental Science, 424, Article 012015. https://doi.org/10.1088/1755-1315/424/1/012015Links ]

Hue, N. (2020). Biochar for maintaining soil health. In B. Giri & A. Varma (Eds.), Soil health (pp. 21-46). Springer International Publishing. https://doi.org/10.1007/978-3-030-44364-1_2Links ]

Huppi, R., Felber, R., Neftel, A., Six, J., & Leifeld, J. (2015). Effect of biochar and liming on soil nitrous oxide emissions from a temperate maize cropping system. Soil, 1(2), 707-717. https://doi.org/10.5194/soil-1-707-2015Links ]

Husna, N., Budianta, D., Munandar, M., & Adipati, N. (2019). Evaluation of several biochar types as inoculant carrier for indigenous phosphate solubilizing microoorganism from acid sulphate soil. Journal of Ecological Engineering, 20(6), 1-8. https://doi.org/10.12911/22998993/109078Links ]

Hyland, C., Hanley, K., Enders, A., Rajkovich, S., & Lehmann, J. (2010, August 1-6). Nitrogen leaching in soil amended with biochars produced at low and high temperatures from various feedstocks [Conference presentation]. 19th world congress of soil science, soil solutions for a changing world. Brisbane, Australia. https://www.iuss.org/19th%20WCSS/Symposium/pdf/0742.pdfLinks ]

International Biochar Initiative. (2015). Standardized product definition and product testing guidelines for biochar that is used in soil. International Biochar Initiative. https://www.biochar-international.org/wp-content/uploads/2018/04/IBI_Biochar_Standards_V2.1_Final.pdfLinks ]

Intani, K., Latif, S., Kabir, A. K. M. R., & Muller, J. (2016). Effect of self-purging pyrolysis on yield of biochar from maize cobs, husks and leaves. Bioresource Technology, 218, 541-551. https://doi.org/10.1016/j.biortech.2016.06.114Links ]

Ippolito, J. A., Laird, D. A., & Busscher, W. J. (2012). Environmental benefits of biochar. Journal of Environment Quality, 41(4), 967-972. https://doi.org/10.2134/jeq2012.0151Links ]

Itskos, G., Nikolopoulos, N., Kourkoumpas, D. -S., Koutsianos, A., Violidakis, I., Drosatos, P., & Grammelis, P. (2016). Energy and the Environment. In S. G. Poulopoulos & V. J. Inglezakis (Eds.), Environment and development, basic principles, human activities, and environmental implications (pp. 363-452). Elsevier. https://doi.org/10.1016/B978-0-444-62733-9.00006-XLinks ]

Jafri, N., Wong, W. Y., Doshi, V., Yoon, L. W., & Cheah, K. H. (2018). A review on production and characterization of biochars for application in direct carbon fuel cells. Process Safety and Environmental Protection, 118, 152-166. https://doi.org/10.1016/j.psep.2018.06.036Links ]

Jeffery, S., Abalos, D., Prodana, M., Bastos, A. C., Van Groenigen, J. W., Hungate, B. A., & Verheijen, F. (2017). Biochar boosts tropical but not temperate crop yields. Environmental Research Letters, 12(5), Article 053001. https://doi.org/10.1088/1748-9326/aa67bdLinks ]

Jenkins, J. R., Viger, M., Arnold, E. C., Harris, Z. M., Ventura, M., Miglietta, F., Girardin, C., Edwards, R. J., Rumpel, C., Fornasi-er, F., Zavalloni, C., Tonon, G., Alberti, G., & Taylor, G. (2017). Biochar alters the soil microbiome and soil function: results of next-generation amplicon sequencing across Europe. Gcb Bioenergy, 9(3), 591-612. https://doi.org/10.1111/gcbb.12371Links ]

Jin, Y., Liang, X., He, M., Liu, Y., Tian, G., & Shi, J. (2016). Manure biochar influence upon soil properties, phosphorus distribution and phosphatase activities: a microcosm incubation study. Chemosphere, 142, 128-135. https://doi.org/10.1016/j.chemosphere.2015.07.015Links ]

Jouhara, H., Ahmad, D., van den Boogaert, I., Katsou, E., Simons, S., & Spencer, N. (2018). Pyrolysis of domestic based feedstock at temperatures up to 300°C. Thermal Science and Engineering Progress, 5, 117-143. https://doi.org/10.1016/j.tsep.2017.11.007Links ]

Juriga, M., & Simansky, V. (2019). Effects of biochar and its reapplication on soil pH and sorption properties of silt loam haplic luvisol. Acta Horticulturae et Regiotecturae, 22(2), 65-70. https://doi.org/10.2478/ahr-2019-0012Links ]

Kammann, C., Ippolito, J., Hagemann, N., Borchard, N., Cayuela, M. L., Estavillo, J. M., Fuertes-Mendizabal, T., Jeffery, S., Kern, J., Novak, J., Rasse, D., Saarnio, S., Schmidt, H. P., Spokas, K., & Wrage-Monnig, N. (2017). Biochar as a tool to reduce the agricultural greenhouse-gas burden - knowns, unknowns and future research needs. Journal of Environmental Engineering and Landscape Management, 25(2), 114-139. https://doi.org/10.3846/16486897.2017.1319375Links ]

Kan, T., Strezov, V., & Evans, T. J. (2016). Lignocellulosic biomass pyrolysis: a review of product properties and effects of pyrolysis parameters. Renewable and Sustainable Energy Reviews, 57, 1126-1140. https://doi.org/10.1016/j.rser.2015.12.185Links ]

Karimi, A., Moezzi, A., Chorom, M., & Enayatizamir, N. (2020). Application of biochar changed the status of nutrients and biological activity in a calcareous soil. Journal of Soil Science and Plant Nutrition, 20, 450-459. https://doi.org/10.1007/s42729-019-00129-5Links ]

Kazemi, R., Ronaghi, A., Yasrebi, J., Ghasemi-Fasaei, R., & Zarei, M. (2019). Effect of shrimp waste-derived biochar and arbuscular mycorrhizal fungus on yield, antioxidant enzymes, and chemical composition of corn under salinity stress. Journal of Soil Science and Plant Nutrition, 19, 758-770. https://doi.org/10.1007/s42729-019-00075-2Links ]

Kolton, M., Harel, Y. M., Pasternak, Z., Graber, E. R., Elad, Y., & Cytryn, E. (2011). Impact of biochar application to soil on the root-associated bacterial community structure of fully developed greenhouse pepper plants. Applied and Environmental Microbiology, 77(14), 4924-4930. https://doi.org/10.1128/AEM.00148-11Links ]

Kwiatkowski, M., & Kalderis, D. (2020). A complementary analysis of the porous structure of biochars obtained from biomass. Carbon Letters, 30, 325-329. https://doi.org/10.1007/s42823-019-00101-4Links ]

Laghari, M., Hu, Z., Mirjat, M. S., Xiao, B., Tagar, A. A., & Hu, M. (2016). Fast pyrolysis biochar from sawdust improves the quality of desert soils and enhances plant growth. Journal of the Science of Food and Agriculture, 96(1), 199-206. https://doi.org/10.1002/jsfa.7082Links ]

Lan, Z. M., Chen, C. R., Rezaei Rashti, M., Yang, H., & Zhang, D. K. (2019). Linking feedstock and application rate of biochars to N2O emission in a sandy loam soil: potential mechanisms. Geoderma, 337, 880-892. https://doi.org/10.1016/j.geoderma.2018.11.007Links ]

Lee, J. W., Hawkins, B., Li, X., & Day, D. M. (2013). Biochar fertilizer for soil amendment and carbon sequestration. In J. W. Lee (Ed.), Advanced biofuels and bioproducts (pp. 57-68). Springer Science and Business Media. https://doi.org/10.1007/978-1-4614-3348-4_6Links ]

Lee, J., Sarmah, A. K., & Kwon, E. E. (2019). Production and formation of biochar. In Y. S. Ok, D. C. W. Tsang, N. Bolan, & J. M. Novak (Eds.), Biochar from biomass and waste, fundamentals and applications (pp. 3-18). Elsevier. https://doi.org/10.1016/B978-0-12-811729-3.00001-7Links ]

Lehmann, J., Gaunt, J., & Rondon, M. (2006). Biochar sequestration in terrestrial ecosystems - A review. Mitigation and Adaptation Strategies for Global Change, 11, 403-427. https://doi.org/10.1007/s11027-005-9006-5Links ]

Lehmann, J., & Joseph, S. (2009). Biochar for environmental management: an Introduction. In J. Lehmann, & S. Joseph (Eds.), Biochar for environmental management, science, technology and implementation (pp. 1-12). Taylor & Francis Group. [ Links ]

Lehmann, J., Rillig, M. C., Thies, J., Masiello, C. A., Hockaday, W. C., & Crowley, D. (2011). Biochar effects on soil biota - A review. Soil Biology and Biochemistry, 43(9), 1812-1836. https://doi.org/10.1016/j.soilbio.2011.04.022Links ]

Leng, L., & Huang, H. (2018). An overview of the effect of pyrolysis process parameters on biochar stability. Bioresource Technology, 270, 627-642. https://doi.org/10.1016/j.biortech.2018.09.030Links ]

Levin, A. A., Shamansky, V. A., & Kozlov, A. N. (2016). A model of pyrolysis in a staged scheme of low-grade solid fuel gasification. Journal of Physics: conference Series, 754, Article 022006. https://doi.org/10.1088/1742-6596/754/2/022006Links ]

Li, X., Chen, X., Weber-Siwirska, M., Cao, J., & Wang, Z. (2018). Effects of rice-husk biochar on sand-based rootzone amendment and creeping bentgrass growth. Urban Forestry and Urban Greening, 35, 165-173. https://doi.org/10.1016/j.ufug.2018.09.001Links ]

Li, Y., Xing, B., Ding, Y., Han, X., & Wang, S. (2020). A critical review of the production and advanced utilization of biochar via selective pyrolysis of lignocellulosic biomass. Bioresource Technology, 312, Article 123614. https://doi.org/10.1016/j.biortech.2020.123614Links ]

Liao, N., Li, Q., Zhang, W., Zhou, G., Ma, L., Min, W., Ye, J., & Hou, Z. (2016). Effects of biochar on soil microbial community composition and activity in drip-irrigated desert soil. European Journal of Soil Biology, 72, 27-34. https://doi.org/10.1016/j.ejsobi.2015.12.008Links ]

Liu, H., Li, H., Zhang, A., Rahaman, M. A., & Yang, Z. (2020). Inhibited effect of biochar application on N2O emissions is amount and time-dependent by regulating denitrification in a wheat-maize rotation system in North China. Science of the Total Environment, 721, Article 137636. https://doi.org/10.1016/j.scitotenv.2020.137636Links ]

Liu, J., Ding, Y., Ma, L., Gao, G., & Wang, Y. (2017). Combination of biochar and immobilized bacteria in cypermethrin-contaminated soil remediation. International Biodeterioration and Biodegradation, 120, 15-20. https://doi.org/10.1016/j.ibiod.2017.01.039Links ]

Liu, N., Sun, Z., Wu, Z., Zhan, X., Zhang, K., Zhao, E., & Han, X. (2013). Adsorption characteristics of ammonium nitrogen by biochar from diverse origins in water. Advanced Materials Research, 664, 305-312. https://doi.org/10.4028/www.scientific.net/AMR.664.305Links ]

Lyu, H., He, Y., Tang, J., Hecker, M., Liu, Q., Jones, P. D., Codling, G., & Giesy, J. P. (2016). Effect of pyrolysis temperature on potential toxicity of biochar if applied to the environment. Environmental Pollution, 218, 1-7. https://doi.org/10.1016/j.envpol.2016.08.014Links ]

Macedo, R. S., Teixeira, W. G., Corrêa, M. M., Martins, G. C., & Vidal-Torrado, P. (2017). Pedogenetic processes in anthrosols with pretic horizon (Amazonian Dark Earth) in Central Amazon, Brazil. PLOS One, 12(5), Article e0178038. https://doi.org/10.1371/journal.pone.0178038Links ]

Martinsen, V., Mulder, J., Shitumbanuma, V., Sparrevik, M., Børresen, T., & Cornelissen, G. (2014). Farmer-led maize biochar trials: effect on crop yield and soil nutrients under conservation farming. Journal of Plant Nutrition and Soil Science, 177(5), 681-695. https://doi.org/10.1002/jpln.201300590Links ]

Masud, M. M., Abdulaha-Al Baquy, M., Akhter, S., Sen, R., Barman, A., & Khatun, M. R. (2020). Liming effects of poultry litter derived biochar on soil acidity amelioration and maize growth. Ecotoxicology and Environmental Safety, 202, Article 110865. https://doi.org/10.1016/j.ecoenv.2020.110865Links ]

Mendes, K. F., Júnior, A. F. D., Takeshita, V., Régo, A. P. J., & Tornisielo, V. L. (2018). Effect of biochar amendments on the sorption and desorption herbicides in agricultural soil. In S. Edebali (Ed.), Advanced sorption process applications (pp. 87-103). IntertechOpen. https://doi.org/10.5772/intechopen.80862Links ]

Muhammad, N., Aziz, R., Brookes, P. C., & Xu, J. (2017). Impact of wheat straw biochar on yield of rice and some properties of Psammaquent and Plinthudult. Journal of Soil Science and Plant Nutrition, 17(3), 808-823. https://doi.org/10.4067/S0718-95162017000300019Links ]

Neina, D. (2019). The role of soil pH in plant nutrition and soil remediation. Applied and Environmental Soil Science, 2019, Article 5794869. https://doi.org/10.1155/2019/5794869Links ]

Nelissen, V., Ruysschaert, G., Manka'Abusi, D., D'Hose, T., De Beuf, K., Al-Barri, B., Cornelis, W., & Boeckx, P. (2015). Impact of a woody biochar on properties of a sandy loam soil and spring barley during a two-year field experiment. European Journal of Agronomy, 62, 65-78. https://doi.org/10.1016/j.eja.2014.09.006Links ]

Nguyen, B. T., Trinh, N. N., & Bach, Q.-V. (2020a). Methane emissions and associated microbial activities from paddy salt-affected soil as influenced by biochar and cow manure addition. Applied Soil Ecology, 152, Article 103531. https://doi.org/10.1016/j.apsoil.2020.103531Links ]

Nguyen, T. K. P., Khoi, C., Ritz, K., Sinh, N., Tarao, M., & Toyota, K. (2020b). Potential use of rice husk biochar and compost to improve P availability and reduce GHG emissions in acid sulfate soil. Agronomy, 10, Article 685. https://doi.org/10.3390/agronomy10050685Links ]

Novak, J. M., Busscher, W. J., Watts, D. W., Laird, D. A., Ahmedna, M. A., & Niandou, M. A. S. (2010). Short-term CO2 mineralization after additions of biochar and switchgrass to a Typic Kandiudult. Geoderma, 154(3-4), 281-288. https://doi.org/10.1016/j.geoderma.2009.10.014Links ]

Odinga, E. S., Waigi, M. G., Gudda, F. O., Wang, J., Yang, B., Hu, X., Li, S., & Gao, Y. (2020). Occurrence, formation, environmental fate and risks of environmentally persistent free radicals in biochars. Environment International, 134, Article 105172. https://doi.org/10.1016/j.envint.2019.105172Links ]

Ogawa, M., & Okimori, Y. (2010). Pioneering works in biochar research, Japan. Australian Journal of Soil Research, 48, 489-500. https://doi.org/10.1071/SR10006Links ]

Oladele, S. O. (2019). Changes in physicochemical properties and quality index of an Alfisol after three years of rice husk biochar amendment in rainfed rice - maize cropping sequence. Geoderma, 353, 359-371. https://doi.org/10.1016/j.geoderma.2019.06.038Links ]

Ouyang, L., Tang, Q., Yu, L., & Zhang, R. (2014). Effects of amendment of different biochars on soil enzyme activities related to carbon mineralisation. Soil Research, 52(7), 706-716. https://doi.org/10.1071/SR14075Links ]

Oyedeji, S., Animasaun, D. A., Ademola, O. I., & Agboola, O. O. (2018). Growth performance of cowpea in spent oil-contaminated soils ameliorated with cocoa shell powder and biochar. Journal of Biological and Environmental Sciences, 12(36), 105-112. [ Links ]

Palansooriya, K. N., Wong, J. T. F., Hashimoto, Y., Huang, L., Rinklebe, J., Chang, S. X., Bolan, N., Wang, H., & Ok, Y. S. (2019). Response of microbial communities to biochar-amended soils: a critical review. Biochar, 1, 3-22. https://doi.org/10.1007/s42773-019-00009-2Links ]

Pariyar, P., Kumari, K., Jain, M. K., & Jadhao, P. S. (2020). Evaluation of change in biochar properties derived from different feedstock and pyrolysis temperature for environmental and agricultural application. Science of the Total Environment, 713, Article 136433. https://doi.org/10.1016/j.scitotenv.2019.136433Links ]

Park, J. H., Ok, Y. S., Kim, S. H., Kang, S. W., Cho, J. S., Heo, J. S., Delaune, R. D., & Seo, D. C. (2015). Characteristics of biochars derived from fruit tree pruning wastes and their effects on lead adsorption. Journal of the Korean Society for Applied Biological Chemistry, 58(5), 751-760. https://doi.org/10.1007/s13765-015-0103-1Links ]

Paustian, K., Lehmann, J., Ogle, S., Reay, D., Robertson, G. P., & Smith, P. (2016). Climate-smart soils. Nature, 532, 49-57. https://doi.org/10.1038/nature17174Links ]

Pratiwi, E. P. A., Hillary, A. K., Fukuda, T., & Shinogi, Y. (2016). The effects of rice husk char on ammonium, nitrate and phosphate retention and leaching in loamy soil. Geoderma, 277, 61-68. https://doi.org/10.1016/j.geoderma.2016.05.006Links ]

Purakayastha, T. J., Bera, T., Bhaduri, D., Sarkar, B., Mandal, S., Wade, P., Kumari, S., Biswas, S., Menon, M., Pathak, H., & Tsang, D. C. W. (2019). A review on biochar modulated soil condition improvements and nutrient dynamics concerning crop yields: pathways to climate change mitigation and global food security. Chemosphere, 227, 345-365. https://doi.org/10.1016/j.chemosphere.2019.03.170Links ]

Rafiq, M. K., Bachmann, R. T., Rafiq, M. T., Shang, Z., Joseph, S., & Long, R. (2016). Influence of pyrolysis temperature on physico-chemical properties of corn stover (Zea mays L.) biochar and feasibility for carbon capture and energy balance. PLOS One, 11(6), Article e0156894. https://doi.org/10.1371/journal.pone.0156894Links ]

Saeid, A., & Chojnacka, K. (2019). Fertlizers: need for new strategies. In S. Chandran, M. R. Unni, & S. Thomas (Eds.), Organic farming, global perspectives and methods (pp. 91-116). Woodhead Publishing. https://doi.org/10.1016/B978-0-12-813272-2.00004-5Links ]

Sahu, P. K., & Brahmaprakash, G. P. (2016). Formulations of bio-fertilizers - approaches and advances. In D. P. Singh, H. B. Singh, & R. Prabha (Eds.), Microbial inoculants in sustainable agricultural productivity (pp. 179-198). Springer. https://doi.org/10.1007/978-81-322-2644-4_12Links ]

Saranya, K., Kumutha, K., & Krishnan, P. S. (2011). Influence of biochar and Azospirillum application on the growth of maize. Madras Agricultural Journal, 98(4/6), 158-164. [ Links ]

Sato, M. K., de Lima, H. V., Costa, A. N., Rodrigues, S., Pedroso, A. J. S., & de Freitas Maia, C. M. B. (2019). Biochar from acai agroindustry waste: study of pyrolysis conditions. Waste Management, 96, 158-167. https://doi.org/10.1016/j.wasman.2019.07.022Links ]

Schimmelpfennig, S., Muller, C., Grunhage, L., Koch, C., & Kammann, C. (2014). Biochar, hydrochar and uncarbonized feedstock application to permanent grassland-Effects on greenhouse gas emissions and plant growth. Agriculture, Ecosystems and Environment, 191, 39-52. https://doi.org/10.1016/j.agee.2014.03.027Links ]

Shaheen, A., & Turaib Ali Bukhari, S. (2018). Potential of sawdust and corn cobs derived biochar to improve soil aggregate stability, water retention, and crop yield of degraded sandy loam soil. Journal of Plant Nutrition, 41(20), 2673-2682. https://doi.org/10.1080/01904167.2018.1509092Links ]

Shahjahan, M., Inamul-Haq, M., Mukhtar, T., & Khalid, A. (2018). Biochar as a carrier of antagonistic rhizobacteria suppressing Macrophomina phaseolina. Transylvanian Review, 26(28), 7469-7476. [ Links ]

Shareef, T. M. E., Zhao, B., & Filonchyk, M. (2018). Characterization of biochars derived from maize straw and corn cob and effects oftheir amendment on maize growth and loess soil properties. Fresenius Environmental Bulletin, 27(5A), 3678-3686. [ Links ]

Sharma, S. P. (2018). Biochar for carbon sequestration: bioengineering for sustainable environment. In D. Barh & V. Azevedo (Eds.), Omics technologies and bio-engineering volume 2: towards improving quality of life (pp. 365-385). Academic Press. https://doi.org/10.1016/B978-0-12-815870-8.00020-6Links ]

Shen, J., Tang, H., Liu, J., Wang, C., Li, Y., Ge, T., Jones, D. L., & Wu, J. (2014). Contrasting effects ofstraw and straw-derived biochar amendments on greenhouse gas emissions within double rice cropping systems. Agriculture, Ecosystems and Environment, 188, 264-274. https://doi.org/10.1016/j.agee.2014.03.002Links ]

Shetty, R., & Prakash, N. B. (2020). Effect of different biochars on acid soil and growth parameters of rice plants under aluminium toxicity. Scientific Reports, 10(1), Article 12249. https://doi.org/10.1038/s41598-020-69262-xLinks ]

Shim, T., Yoo, J., Ryu, C., Park, Y. K., & Jung, J. (2015). Effect of steam activation ofbiochar produced from a giant Miscanthus on copper sorption and toxicity. Bioresource Technology, 197, 85-90. https://doi.org/10.1016/j.biortech.2015.08.055Links ]

Sigmund, G., Huffer, T., Hofmann, T., & Kah, M. (2017). Biochar total surface area and total pore volume determined by N2 and CO2 physisorption are strongly influenced by degassing temperature. Science of the Total Environment, 580, 770-775. https://doi.org/10.1016/j.scitotenv.2016.12.023Links ]

Smith, P. (2016). Soil carbon sequestration and biochar as negative emission technologies. Global Change Biology, 22(3), 1315-1324. https://doi.org/10.1111/gcb.13178Links ]

Spokas, K. A., Baker, J. M., & Reicosky, D. C. (2010). Ethylene: potential key for biochar amendment impacts. Plant and Soil, 333(1-2), 443-452. https://doi.org/10.1007/s11104-010-0359-5Links ]

Stewart, C. E., Zheng, J., Botte, J., & Cotrufo, M. F. (2013). Co-generated fast pyrolysis biochar mitigates green-house gas emissions and increases carbon sequestration in temperate soils. GCB Bioenergy, 5(2), 153-164. https://doi.org/10.1111/gcbb.12001Links ]

Sun, H., Zhang, H., Shi, W., Zhou, M., & Ma, X. (2019). Effect of biochar on nitrogen use efficiency, grain yield and amino acid content of wheat cultivated on saline soil. Plant, Soil and Environment, 65(2), 83-89. https://doi.org/10.17221/525/2018-PSELinks ]

Szwaja, S., Poskart, A., & Zajemska, M. (2018, June 26-29). A new approach for evaluating biochar quality from biomass thermal processing [Conference presentation]. 3rd International Conference on Smart and Sustainable Technologies (SpliTech), Split, Croatia. https://ieeexplore.ieee.org/document/8448316Links ]

Tabakaev, R., Kanipa, I., Astafev, A., Dubinin, Y., Yazykov, N., Zavorin, A., & Yakovlev, V. (2019). Thermal enrichment of different types of biomass by low-temperature pyrolysis. Fuel, 245, 29-38. https://doi.org/10.1016/j.fuel.2019.02.049Links ]

Tag, A. T., Duman, G., Ucar, S., & Yanik, J. (2016). Effects of feedstock type and pyrolysis temperature on potential applications of biochar. Journal of Analytical and Applied Pyrolysis, 120, 200-206. https://doi.org/10.1016/j.jaap.2016.05.006Links ]

Taghizadeh-Toosi, A., Clough, T. J., Condron, L. M., Sherlock, R. R., Anderson, C. R., & Craigie, R. A. (2011). Biochar incorporation into pasture soil suppresses in situ nitrous oxide emissions from ruminant urine patches. Journal of Environment Quality, 40(2), 468-476. https://doi.org/10.2134/jeq2010.0419Links ]

Tan, Z., Lin, C. S. K., Ji, X., & Rainey, T. J. (2017). Returning biochar to fields: a review. Applied Soil Ecology, 116, 1-11. https://doi.org/10.1016/j.apsoil.2017.03.017Links ]

Tomczyk, A., Sokolowska, Z., & Boguta, P. (2020). Biochar physicochemical properties: pyrolysis temperature and feedstock kind effects. Reviews in Environmental Science and Bio/Technology, 19, 191-215. https://doi.org/10.1007/s11157-020-09523-3Links ]

Uddin, M. N., Techato, K., Taweekun, J., Rahman, M. M., Rasul, M. G., Mahlia, T. M. I., & Ashrafur, S. M. (2018). An overview of recent developments in biomass pyrolysis technologies. Energies, 11(11), Article 3115. https://doi.org/10.3390/en11113115Links ]

Vecstaudza, D., Senkovs, M., Nikolajeva, V., Kasparinskis, R., & Muter, O. (2017). Wooden biochar as a carrier for endophytic isolates. Rhizosphere, 3, 126-127. https://doi.org/10.1016/j.rhisph.2017.04.002Links ]

Ventura, M., Alberti, G., Panzacchi, P., Vedove, G. D., Miglietta, F., & Tonon, G. (2019). Biochar mineralization and priming effect in a poplar short rotation coppice from a 3-year field experiment. Biology and Fertility of Soils, 55, 67-78. https://doi.org/10.1007/s00374-018-1329-yLinks ]

Wang, X., Li, X., Liu, G., He, Y., Chen, C., Liu, X., Li, G., Gu, Y., & Zhao, Y. (2019). Mixed heavy metal removal from wastewater by using discarded mushroom-stick biochar: adsorption properties and mechanisms. Environmental Science: processes and Impacts, 21(3), 584-592. https://doi.org/10.1039/C8EM00457ALinks ]

Wang, Y., Yin, R., & Liu, R. (2014). Characterization of biochar from fast pyrolysis and its effect on chemical properties of the tea garden soil. Journal of Analytical and Applied Pyrolysis, 110, 375-381. https://doi.org/10.1016/j.jaap.2014.10.006Links ]

Weber, K., & Quicker, P. (2018). Properties of biochar. Fuel, 217, 240-261. https://doi.org/10.1016/j.fuel.2017.12.054Links ]

Woolf, D., Amonette, J. E., Street-Perrott, F. A., Lehmann, J., & Joseph, S. (2010). Sustainable biochar to mitigate global climate change. Nature Communications, 1, Article 56. https://doi.org/10.1038/ncomms1053Links ]

Wu, Q., Xian, Y., He, Z., Zhang, Q., Wu, J., Yang, G., Zhang, X., Qi, H., Ma, J., Xiao, Y., & Long, L. (2019). Adsorption characteristics of Pb(II) using biochar derived from spent mushroom substrate. Scientific Reports , 9(1), Article 15999. https://doi.org/10.1038/s41598-019-52554-2Links ]

Xiao, R., Yang, W., Cong, X., Dong, K., Xu, J., Wang, D., & Yang, X. (2020). Thermogravimetric analysis and reaction kinetics of lignocellulosic biomass pyrolysis. Energy, 201, Article 117537. https://doi.org/10.1016/j.energy.2020.117537Links ]

Yaashikaa, P. R., Senthil Kumar, P., Varjani, S. J., & Saravanan, A. (2019). Advances in production and application of biochar from lignocellulosic feedstocks for remediation of environmental pollutants. Bioresource Technology, 292, Article 122030. https://doi.org/10.1016/j.biortech.2019.122030Links ]

Yadav, V., Karak, T., Singh, S., Singh, A. K., & Khare, P. (2019). Benefits of biochar over other organic amendments: responses for plant productivity (Pelargonium graveolens L.) and nitrogen and phosphorus losses. Industrial Crops and Products, 131, 96-105. https://doi.org/10.1016/j.indcrop.2019.01.045Links ]

Yang, H., Yan, R., Chen, H., Lee, D. H., & Zheng, C. (2007). Characteristics of hemicellulose, cellulose and lignin pyrolysis. Fuel, 86, 1781-1788. https://doi.org/10.1016/j.fuel.2006.12.013Links ]

Yao, Y., Gao, B., Zhang, M., Inyang, M., & Zimmerman, A. R. (2012). Effect of biochar amendment on sorption and leaching of nitrate, ammonium, and phosphate in a sandy soil. Chemosphere, 89(11), 1467-1471. https://doi.org/10.1016/j.chemosphere.2012.06.002Links ]

Ye, L., Camps-Arbestain, M., Shen, Q., Lehmann, J., Singh, B., & Sabir, M. (2020). Biochar effects on crop yields with and without fertilizer: a meta-analysis of field studies using separate controls. Soil Use and Management, 36, 2-18. https://doi.org/10.1111/sum.12546Links ]

Yu, H., Zou, W., Chen, J., Chen, H., Yu, Z., Huang, J., Tang, H., Wei, X., & Gao, B. (2019). Biochar amendment improves crop production in problem soils: a review. Journal of Environmental Management, 232, 8-21. https://doi.org/10.1016/j.jenvman.2018.10.117Links ]

Yu, J., Paterson, N., Blamey, J., & Millan, M. (2017). Cellulose, xylan and lignin interactions during pyrolysis of lignocel-lulosic biomass. Fuel, 191, 140-149. https://doi.org/10.1016/j.fuel.2016.11.057Links ]

Zhang, C., Liu, L., Zhao, M., Rong, H., & Xu, Y. (2018). The environmental characteristics and applications of biochar. Environmental Science and Pollution Research, 25(22), 21525-21534. https://doi.org/10.1007/s11356-018-2521-1Links ]

Zhang, C., Zeng, G., Huang, D., Lai, C., Chen, M., Cheng, M., Tang, W., Tang, L., Dong, H., Huang, B., Tan, X., & Wang, R. (2019). Biochar for environmental management: mitigating greenhouse gas emissions, contaminant treatment, and potential negative impacts. Chemical Engineering Journal, 373, 902-922. https://doi.org/10.1016/j.cej.2019.05.139Links ]

Zhang, H., Chen, C., Gray, E. M., & Boyd, S. E. (2017). Effect of feedstock and pyrolysis temperature on properties of biochar governing end use efficacy. Biomass and Bioenergy, 105, 136-146. https://doi.org/10.1016/j.biombioe.2017.06.024Links ]

Zhao, J., Shen, X. J., Domene, X., Alcaniz, J. M., Liao, X., & Palet, C. (2019). Comparison of biochars derived from different types of feedstock and their potential for heavy metal removal in multiple-metal solutions. Scientific Reports , 9, Article 9869. https://doi.org/10.1038/s41598-019-46234-4Links ]

Zhao, L., Cao, X., Masek, O., & Zimmerman, A. (2013). Heterogeneity of biochar properties as a function of feedstock sources and production temperatures. Journal of Hazardous Materials, 256-257, 1-9. https://doi.org/10.1016/j.jhazmat.2013.04.015Links ]

Zhou, Z., Shi, D., Qiu, Y., & Sheng, G. D. (2010). Sorptive domains of pine chars as probed by benzene and nitrobenzene. Environmental Pollution, 158, 201-206. https://doi.org/10.1016/j.envpol.2009.07.020Links ]

Zimmerman, A. R. (2010). Abiotic and microbial oxidation of laboratory-produced black carbon (biochar). Environmental Science and Technology, 44(4), 1295-1301. https://doi.org/10.1021/es903140cLinks ]

Zong, Y., Wang, Y., Sheng, Y., Wu, C., & Lu, S. (2018). Ameliorating soil acidity and physical properties of two contrasting texture Ultisols with wastewater sludge biochar. Environmental Science and Pollution Research, 25, 25726-25733. https://doi.org/10.1007/s11356-017-9509-0Links ]

Zwart, K. (2020). Effects of biochar produced from waste on soil quality. In E. Meers, G. Velthof, E. Michels, & R. Rietra (Eds.), Biorefinery of Inorganics: recovering mineral nutrients from biomass and organic waste (pp. 283-299). John Wiley & Sons Ltd. https://doi.org/10.1002/9781118921487.ch5-7Links ]

Author's contributions SMR carried out the writing of the original draft and TGH conducted the critical review and revision of this manuscript.

Received: May 18, 2020; Accepted: December 11, 2020

*Corresponding author: tghneim@icesi.edu.co

Creative Commons License This is an open-access article distributed under the terms of the Creative Commons Attribution License