Download as pdf or txt
Download as pdf or txt
You are on page 1of 54

Critical Reviews in Environmental Science and

Technology

ISSN: 1064-3389 (Print) 1547-6537 (Online) Journal homepage: https://www.tandfonline.com/loi/best20

Progress and future prospects in biochar


composites: Application and reflection in the soil
environment

Sandip Mandal, Shengyan Pu, Sangeeta Adhikari, Hui Ma, Do-Heyoung Kim,
Yingchen Bai & Deyi Hou

To cite this article: Sandip Mandal, Shengyan Pu, Sangeeta Adhikari, Hui Ma, Do-Heyoung Kim,
Yingchen Bai & Deyi Hou (2020): Progress and future prospects in biochar composites: Application
and reflection in the soil environment, Critical Reviews in Environmental Science and Technology,
DOI: 10.1080/10643389.2020.1713030

To link to this article: https://doi.org/10.1080/10643389.2020.1713030

View supplementary material

Published online: 24 Jan 2020.

Submit your article to this journal

View related articles

Full Terms & Conditions of access and use can be found at


https://www.tandfonline.com/action/journalInformation?journalCode=best20
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY
https://doi.org/10.1080/10643389.2020.1713030

Progress and future prospects in biochar composites:


Application and reflection in the soil environment
Sandip Mandal a, Shengyan Pua,b, Sangeeta Adhikaric, Hui Maa,d,
Do-Heyoung Kimc, Yingchen Baib, and Deyi Houe
a
State Key Laboratory of Geohazard Prevention and Geoenvironment Protection, Chengdu
University of Technology, Chengdu, P. R. China; bState Key Laboratory of Environmental Criteria
and Risk Assessment, Chinese Research Academy of Environmental Sciences, Beijing, P. R. China;
c
School of Chemical Engineering, Chonnam National University, Gwangju, Korea; dDepartment of
Plant and Environmental Sciences, University of Copenhagen, Frederiksberg, Denmark; eSchool of
Environment, Tsinghua University, Beijing, P. R. China

ABSTRACT
In recent years, environmental
pollution is a major global con-
cern and adversely affecting
human health and the economy.
The environmental properties of
soil are exceptionally complex,
which makes the remediation
and treatment more challenging
and expensive. In this aspect, sus-
tainable materials (biomaterials,
biochar, and composites) present
an effective and efficient remediation solution. Exactly, biochar and its relative compo-
sites can achieve novel properties with addition of metal oxides, surface agents, and
nano-material. The present review study provides quantitative analysis and addresses
the advantages of using biochar composites relative with pure biochar. The synthesis
of Nano-metal assisted biochar, properties and utilization in improving soil properties
as well as heavy metal removal are reviewed. Specific attention has been paid to the
impacts of biochar and its composite on soil fundamental properties such as pH, cation
exchange efficiency, bulk density, porosity, water retention capability, soil organic mat-
ter (SOM) and redox reactions. Significant emphasis is focused on soil remediation, in
particular, the elimination of heavy metals utilizing biochar-based composites. Further
research is urged to verify the proposed mechanisms involved biochar composites-
microbial interactions and heavy metal remediation considering short term and long
term efficiency for improvement.

HIGHLIGHTS
 Progress in development, synthesis methods, and application of biochar-based
composites in the soil are studied.
 The impact mechanisms of biochar composites on essential soil properties have
been studied.
 Biochar-based composites have a significant effect on bulk density and water reten-
tion capacity.

CONTACT Shengyan Pu pushengyan@gmail.com, pushengyan13@cdut.edu.cn State Key Laboratory of


Geohazard Prevention and Geoenvironment Protection, Chengdu University of Technology, Chengdu 610059,
P. R. China.
Supplemental data for this article is available online at https://doi.org/10.1080/10643389.2020.1713030
ß 2020 Taylor & Francis Group, LLC
2 S. MANDAL ET AL.

 Biochar-based composites were found to be better than pristine biochar in soil


heavy metal remediation.
 The potential use of biochar-based composites in different soil heavy metal remedi-
ation prospected.

KEYWORDS Biochar composites; soil heavy metals; soil pollution remediation; nano-metal assisted biochar;
immobilization

1. Introduction
The biochar is a growing interest in accepting carbon potential and applica-
tions, especially its environmental management and mitigation of carbon
dioxide from the environment. In fact, studies have identified biochar as an
excellent soil additives that can change the perception of environmental
management (Enders, Hanley, Whitman, Joseph, & Lehmann, 2012;
Lehmann & Joseph, 2012a, 2012b; Whitman & Lehmann, 2009). Interest in
the investigation has gradually increased, given the multi-application of
biochar related to immobilization (Park, Choppala, Bolan, Chung, &
Chuasavathi, 2011; Uchimiya, Lima, Klasson, & Wartelle, 2010; Uchimiya,
Lima, Thomas Klasson, et al., 2010), soil redox management (Zhu, Chen,
Zhu, & Xing, 2017), soil fertility (Ok, Uchimiya, Chang, & Bolan, 2016),
carbon sequestration (Lehmann et al., 2011), aggregates stabilization (Gale,
Cambardella, & Bailey, 2000), and greenhouse gas reduction (Gaunt &
Cowie, 2012). The effects of biochar on soil nutrient recycling, water reten-
tion, soil fertility, and other important soil properties need to be studied
further (Zhu et al., 2017). Biochar’s chemical composition depends upon
the type of raw materials and pyrolysis conditions such as pyrolysis temper-
atures, reaction intervals, reaction atmosphere (oxidative/reductive), and
firing rate. Thus, not all biochar are equal, and it is difficult to determine
the exact chemical composition of biochar (Lehmann et al., 2011). In add-
ition, the procedure of carbonization at high temperatures (>200  C) and
conversion of biomass to biochar can generate biofuels and syngas as useful
by-products (Lehmann & Joseph, 2012a).
Molecular analysis of biochar confirmed that after heat treatment (400  C
or above) of plant biomass will undergo chemical changes and attained a
more stable molecular structure. Elemental, thermal analysis, solid-state 13C
Nuclear magnetic resonance (NMR), and Fourier transform infrared (FT-
IR) spectra show that, starting from the temperature of 255  C, most of dry
biomass of plants including cellulose, pectin and lignin, are mainly altered
by dehydration to phenol, furan, aromatic and some alkyl carbon struc-
tures, with a large amount of O, H, and S (Qian et al., 2019; Tan, Lin, Ji, &
Rainey, 2017). In particular, the biochar developed at relatively low
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY 3

temperatures 200–400  C has more oxygen-containing functional groups,


such as -COOH,-OH, C ¼ O, phenolic-OH and-CHO groups, thus improv-
ing the immobilization and binding efficiency of inorganic pollutants in
the soil.
Additionally, the oxygen-containing functional groups have exhibited the
hydrophilic, hydrophobic and buffering ability to the acidic and alkaline
flux in the soil with higher ion exchange capacities (Sohi, Krull, Lopez-
Capel, & Bol, 2010; Yao, Gao, Zhang, Inyang, & Zimmerman, 2012;
Zimmerman, Gao, & Ahn, 2011). At 400–500  C and higher temperatures,
progressive polymerization, functional groups loss, dehydrogenation, a
decrease in ion-exchange capacity, pores closure and removal of substituted
bases occur, resulting in larger fused aromatic rings. These concentrated
structures are likely to be highly resistant to biological activities in soil
(Keiluweit, Nico, Johnson, & Kleber, 2010). Thus, the concept of “pyrolysis”
is attributed to the degree and extent of condensation in the chemical
structure represented by biochar, the increase in temperature from slightly
charred biomass to soot, and the implicit environmental range of persist-
ence. Over the years, many researchers have reported similar observations
(Ladygina & Rineau, 2013; Ok, Uchimiya, Chang, & Bolan, 2016) and these
challenges could be interpreted and enhanced by the introduction of metal
oxides/Nano-metals/hybrid metal-organic frameworks.
In comparison to pristine biochar, biochar nanocomposites emerged as
novel and promising materials for soil properties improvement, similarly to
carbon mitigation, nutrient enrichment, microbial activity, and toxic heavy
metals removal (Oliveira et al., 2017; Qian, Kumar, Zhang, Bellmer, &
Huhnke, 2015; Qian et al., 2019; Yuan et al., 2017). The study also points out
the inherent features of lignocellulose and the pyrolysis temperature of the
medium, and raw biochar prepared directly from the biomass system usually
have limited surface functional groups and lower surface area (Cho et al.,
2019; Tan et al., 2017). In addition, separation and reuse are challenging when
powder biochar is used for heavy metals removal (Ye, Zeng, Wu, Zhang, Dai,
et al., 2017). Therefore, with high specific surface area, superior reactivity, and
magnetism could overcome the inherent shortcomings and improve its feasi-
bility for various application in soil and water systems (Ho, Zhu, & Chang,
2017; Lyu et al., 2018; Xie et al., 2017; Yuan et al., 2017). Using hydrophilic
flexibility and hierarchical pores of maize stalk biochar-supported zero-valent
iron composite (Fe0-HCS), the benefits of iron nanoparticles with porous bio-
char were obtained (Yang, Zhang, et al., 2018). Similarly, other reports for
developing nZVI based biochar composite are well documented (Wang, Chen,
et al., 2019; Xiong et al., 2017; Yu, Xiong, et al., 2019). Considering this, it
can be said that, the production of biochar-based nanomaterials can achieve
seven integrated objectives, including pollutant removal (Ye, Zeng, Wu,
4 S. MANDAL ET AL.

Zhang, Liang, et al., 2017), waste management, carbon sequestration, energy


production (Chen, Yu, et al., 2018; Yang, Yu, et al., 2019), agricultural produc-
tion, soil physio-chemical and biological process (Patil, Shedbalkar,
Truskewycz, Chopade, & Ball, 2016; Sun et al., 2019).
The preparation of biochar composites can be engineered by two treatment
processes: (1) pre-treatment of biomass and (2) post-treatment of biochar,
which are reviewed and discussed in the following sections. However, in add-
ition to its advantages, it is necessary to consider the environmental and bio-
logical toxicity of nanomaterials (Xie et al., 2017; Yang, Zhang, et al., 2019).
So far, many papers refer to the relative problems of biochar were docu-
mented (Abbas et al., 2018; Cha et al., 2016; Lehmann et al., 2011; Mohan,
Sarswat, Ok, & Pittman, 2014; Patra, Panda, & Dhal, 2017; Premarathna et al.,
2019; Sajjadi, Chen, & Egiebor, 2019; Sharma, Singh, Baskar, & Kumar, 2018;
Sohi et al., 2010; Tan et al., 2015; Vijayaraghavan, 2016; Wang, Zhao, et al.,
2019; Wu et al., 2017; Yaashikaa, Senthil Kumar, Varjani, & Saravanan, 2019;
Zama et al., 2018; Zhang, Zeng, et al., 2019; Zhang, Wang, et al., 2013). Most
of these review papers focused on the preparation or alteration of the pollu-
tants or on their effects in agriculture, soil quality, or the sequestration of car-
bon have been studied and quantitative analysis, were presented. Some of the
documents are restricted to pollutants on water treatment, and the research
about a very high extent soil-based pollutants are significantly missing. The
application of biochar based composites in soil and countereffect and interac-
tions with plant species, microbial community, and soil animals are not well
discussed. Hence, we focused on the new summarized report collecting infor-
mation and studies, which mainly reported biochar composites for heavy
metal remediation in the soil and its impact on soil fundamental properties
during treatments. Furthermore, a comparative evaluation compared with bare
biochar has been represented. The uncertainties and future research that has
been discussed for scale up, and application of biochar-based composites is
also discussed.

2. Biochar composites
Synthesis of biochar-based composites offer advantages for biochar and nano-
materials. The resulting composites usually have significant improvement of
functional groups, pore properties, surface activity sites, catalytic degradation
capacity, ease of separation, etc.

2.1. Production practices


Biochar is the solid carbonaceous product obtained after the controlled pyr-
olysis of biomass (Lyu, He, et al., 2016). In addition to pyrolysis, biochar
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY 5

can also be generated by other techniques such as hydrothermal carboniza-


tion and gasification. Initially, the process starts after feedstocks and waste
materials identified and carefully measured with screening, cleaning and
finally air drying. The raw materials and its physiochemical properties play
an essential role in the development and further production of biochar
composites. The study reported for biochar based composites obtained
from feedstocks such as pine (Mandal, Bhattacharya, Verma, & Haydary,
2018), oak (Aung et al., 2018), oak bark (Mohan, Kumar, Sarswat,
Alexandre-Franco, & Pittman, 2014), rapeseed (Salam et al., 2019), sugar-
cane bagasse (Yi, Tu, Zhao, Tsang, & Fang, 2019), maize (Ralebitso & Orr,
2016), wheat straw (Gła˛b, Palmowska, Zaleski, & Gondek, 2016), bamboo
(Christianson, Hedley, Camps, Free, & Saggar, 2010), and peanut shells (Yi
et al., 2019).
Potential feedstocks for biochar production, either single constituent or
as mixtures, are almost infinite. Possibly, all wood and non-wood based
biomass can be transformed into biochar. The practical application of the
feedstock often has minimal moisture or mineral content depending on the
equipment used. For example, chlorine and alkali metals (potassium, K)
may cause problems with oxidation and high ash content in the resulting
biochar materials (Steiner, Das, Garcia, F€ orster, & Zech, 2008). The proper-
ties of the biochar produced can vary widely as a result of different produc-
tion technologies and feedstocks. The characteristics of the feedstock are
considered to be most significant. During thermal degradation, most of the
fundamental elements such as Carbon (C), Hydrogen (H), Oxygen (O),
Nitrogen (N), and Sulfur (S) are easily volatilized, and elements such as
Calcium (Ca), Magnesium (Mg), Iron (Fe), Manganese (Mn), Silicon (Si),
and Phosphorus (P) remains in the resulting biochar. As a result, biochar
produced with a higher mineral content has a high ash content and this
ash content is mainly responsible for alkalinity in developed biochar, which
is considered to be significant for soil (Gaskin, Steiner, Harris, Das, &
Bibens, 2008). The effect of the feedstock on the development of biochar
determines the quantity and quality of carbon remaining in the biochar.
Through separating the minerals (ash content) from the feedstock and bio-
char from the total mass, it is possible to report the total carbon content.
The carbon is composed of relatively stable aromatic carbon and relatively
volatile aliphatic carbon. With increasing thermal decomposition tempera-
ture, the amount of aliphatic carbon reduces. For instance, it is reported by
Sun et al. (2014) that higher production levels could be obtained at lower
temperatures for biochar and hydro chars. By contrast, higher pyrolysis
temperatures not only increased the content of biochar carbon but also
increased thermal stability. Additionally, inclusive of nanomaterial and
metal oxides to biochar could also increase thermal stability (Li et al.,
6 S. MANDAL ET AL.

2017). According to the above studies by Sun et al., three different types of
biomass were used as feedstock materials: hickory wood (HW), bagasse
(BG), and bamboo (BB), and slow pyrolysis were considered at three differ-
ent temperatures (300, 450, and 600  C). The higher production rate was
observed at a lower temperature. However, biochar yields decrease with the
temperature increased, which is mainly reasoned for moisture and hydra-
tion. In addition, the decomposition of organic matter high-temperature is
also observed. Carbon content rise for bamboo biochar and oxygen content
fell with increased temperatures; while the other two biochar samples (hick-
ory wood and bagasse) reached their peak carbon and oxygen contents at
450  C and then remained stable. For all biochar samples, and particularly
for three hydrochar samples N, P, K, Ca, Mg, Cu, Fe, and Al showed low
values (less than 1%) (Sun et al., 2014). Similarly, three different feedstocks
(wood, agriculture and industrial sewage sludge wastes) were examined for
heavy metal removals (Cr (III), Cd (II), Cu (II) and Pb (II) ions). The
results show that the ability to adsorb depends on the feedstock type and
pyrolysis. The adsorption capacity of the biomass types is ordered as fol-
lows: sewage sludge > Brassica napus (LO) biomasses > corn (ZO) > pop-
lar biomass (CO, from wood). The study clearly indicates that biochars
from different feedstocks (mainly biochar derived from wood and agricul-
tural biomasses) improve the adsorption efficiency. The reason ascertained
that the pore structure is a key factor that can influence the sorption of
heavy metals onto biomass (High porosity and large pore size). In addition,
more active groups (carboxylic groups) are advantageous and differ by
feedstock (Zhao et al. 2019). In a similar studies by Sohi et al. (Sohi et al.,
2010) has also been considered, that different feedstocks have resulted in
different surface area, pores and functional groups in biochar, all of which
have an impact on the sorption characteristics of biochar. Sun et al. (2011)
reported that, despite the two biochar (poultry-litter and weed-straw bio-
char) obtained under the same temperature (400  C) poultry-litter have a
wider specified surface area and porosity than weed-straw biochar.
The pyrolysis techniques such as rapid pyrolysis, slow pyrolysis, and flash
pyrolysis could be considered for biochar composites preparation
(Lehmann & Joseph, 2012b; Weber & Quicker, 2018). Rapid pyrolysis is
categorized by a fast heating speed (greater than 200 K/min) at medium
temperature (400–600  C) and a short residence time (2 seconds). The
process is more appropriate for the production of liquid and natural gas
products. Slow pyrolysis is characterized by slow heating (5–10  C), longer
steam residence times (minutes to hours) and a wider temperature range
(300–800  C). Typically slow and controlled pyrolysis offered better biochar
with advantages of uniform pores and high surface area (Zhao et al., 2018).
The residence time and heating rate will also affect the biochar formation,
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY 7

and the introduction of metal components to biochar can improve its per-
formance (Lian & Xing, 2017).
The physicochemical properties of biochar composites, such as surface
area, water holding capacity, permeability, conductivity, ion-exchange cap-
acity, pH and porosity are significantly varied with pyrolysis conditions. In
the process of increasing the temperature (from 200 to 800  C), the pore
structure and improvement of the biochar surface can be gradually
improved by removing simple molecules such as ethylene and esters from
the outer surface of the raw material. However, the high-temperature flux
may also cause deformation and collapse of some microporous structures.
In addition to the surface area, the reduction of the surface oxy group, and
the increase of temperature can also improve the alkalinity value of the bio-
char due to the crystallization of the inherent minerals. Yao et al. (2011a)
observed that the size of the minimum aromatic fusion ring cluster by
NMR, and the carbon content increased from high temperature. In another
aspect, due to the nature of raw materials, the different components of bio-
mass components will significantly affect its decomposition mechanism and
biochar performance. The composites have the advantage of using blocks
of size in the nano-meter range to design and manufacture new materials
with unprecedented flexibility and physical properties (Hansen et al., 2016).
This carbon-rich material with nanomaterials presented as multifunctional
material and could be produced by simple techniques and chemical modifi-
cations (Ahmed, Zhou, Ngo, Guo, & Chen, 2016). The two common
approaches are widely employed to prepare biochar based composites
(Figure 1): 1) pre-treatment of biomass with Nano agents/metal oxides/
hydroxides, 2) post-treatment of obtained biochar using metal oxides/metal
oxides/hydroxides.

2.1.1. Pre-treatment of biomass with nano metal oxides/metal-


oxides/hydroxides
In pretreatment synthesis techniques, the Nano-metal oxides/metal oxides/
hydroxides/chemical agents were commonly doped to biomass before pyr-
olysis. For example, enrichment of Mg into tomato tissue and slow pyroly-
sis to concentrate the Mg in biochar (Yao et al., 2013). In this work,
tomato plants treated with a small amount of Mg have accumulated high
magnesium content in tissues, which indicates that magnesium tomato
plants can be enriched in large quantities, and the pyrolysis process contin-
ues to concentrate magnesium (8.8% magnesium) in the biochar. The
resulting magnesium doped biochar composites (MgEC) have good phos-
phate adsorption performance in aqueous solutions. The methodology is
bioaccumulation before pyrolysis to enrich metal-based biochar’s and con-
verting into biochar composites under N2 surroundings at temperatures of
8 S. MANDAL ET AL.

Figure 1. Two common approach for preparation of biochar based composites.

600  C for 1 h. The micro-imaging technology reveals the wide distribution


of Nano-magnesium flakes (Nano-platelets) on the carbon surface, which is
a collective structural form of magnesium hydroxide Nano-slice Mg(OH) 2
and MgO granules, which is the most important adsorption position for
phosphate on the carbon surface of Mg-biochar surface. The presence of a
large number of nanoparticles may lead to a high surface area of biochar.
In general, saturated biomass can be transformed into biochar with better
properties through carbonization.
In another study, biochar-Al-O-OH nanocomposites have been produced
through pyrolysis of AlCl3 pretreated with biomass. The Al nanoparticles
loaded on the surface of biochar have significantly increased the reaction
area and activity sites (Figure 2). Nano-Al-OOH are embedded in the bio-
char matrix, which prevents the accumulation of evenly distributed sub-
stances on the surface of biochar (Zhang & Gao, 2013).
Superior property and higher removal capacity for heavy metals and
toxic pollutants from soil and water environment were observed because of
alteration of surface properties. However, after a long time period, the lon-
gevity and effectiveness of the biochar composite decreased, and the reason
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY 9

Figure 2. (a) TEM images of biochar/AlOOH; (b) EDX analysis of selected areas marked with
blue markings (c) EDX analysis of biochar/AlOOH. The inserts in (a) show the selection of
regional electron diffraction (SAED) patterns. (To explain the reference to color in this figure
legend, refer to the Web version of this article) (Zhang & Gao, 2013).

ascertained to complexation with several ions available during treatment,


oxidation of the reactive metal component, aggregation of reactive metal in
the biochar surface, microbial digestion and breakdown of the carbon
structure. To overcome these difficulties, emphasis was given to develop
sustainable materials with functional Nano-particles, which can be incorpo-
rated into the biochar to extend its surface area and increase its high affin-
ity, thermal stability and/or cation exchange strength (Wan et al., 2019).
For the purpose of magnetic- biochar is well studied, explored, and utilized
for various environmental applications. Incorporating functional nanopar-
ticles into the biochar can expand its surface area and increase its high-
affinity adsorption sites, thermal stability, and/or cation exchange capacity.
In similar fashion as stated above, the nZVI was introduced to biochar sur-
face to make the material magnetic. For example, biochar-c-Fe2O3 compo-
sites were produced by Nano-sized c-Fe2O3 particles embedded in porous
biochar matrix by thermal pyrolysis of the biomass (600  C/1 h).
The obtained microscopy images show that Nano and micro c-Fe2O3
particles are dispersed on the surface of biochar without aggregation
(Figure 3). The c-Fe2O3 particles are partially embedded in the biochar
matrix, which indicates that there is a good mechanical binding between
the biochar matrix and the iron oxide particles, thus preventing the
c-Fe2O3 particles being separated from the biochar matrix caused by unre-
lated forces (Zhang, Gao, et al., 2013). In a similar approach for the pro-
duction of biochar with metal and oxides are reported by the authors
10 S. MANDAL ET AL.

Figure 3. (i) Depiction of the preparation strategy of biochar/c-Fe2O3 composite. (ii) SEM image
of biochar/c-Fe2O3 composite: (a) and (b) morphological structures; (c) Micro-sized c-Fe2O3 par-
ticles; and (d) nano-c-Fe2O3 particles (Zhang, Gao, et al., 2013).

(Agrafioti, Kalderis, & Diamadopoulos, 2014; Han et al., 2015; Lee, Kim,
Choi, Chung, & Mahendran, 2014; Peng & Sun, 2010; Saadat, Raei, &
Talebbeydokhti, 2018). In summary, the heat treatment of biological carbon
and metal oxide solution mixtures and the formation of oxygen-containing
functional groups of surface complexes are the key factors to improve the
efficiency of metal oxides/Nano-metal oxides supported biochar for envir-
onmental applications. In this regard, further research and investigation are
still required.

2.1.2. Post-treatment of biochar with nano metal oxides/metal


oxides/hydroxides
The impregnation of metal oxide/nanoparticles to pristine biochar could
provide better properties after pyrolysis. Evaporation, heat treatment, con-
ventional wet impregnation, and direct hydrolysis are common methods
for the treatment of biochar in the presence of metallic salts (Tan et al.,
2016). These metal oxide nanoparticles could enhance the surface area,
chemical reactivity and serve as active agents for binding with pollutants in
soil and water. For instance, recently Zhu et al. study modified magnetic
biochar, the approach is a similar trend of impregnation and used to mod-
ify the biochar surface properties. Similarly, Devi and Saroha (Devi &
Saroha, 2014), has developed ZVI-MBC from paper mill sludge by pyrolysis
at 700  C. In the synthesis process of ZVI-MBC, Fe (II) was reduced to Fe
(0) using sodium borohydride (NaBH4) as a reducing agent, and cetyltrime-
thylammonium bromide (CTMB) was added as the surfactant/stabilizer.
Initially, the biochar– CTMB complex was prepared and after complexity,
the biochar– CTMB complex was magnetized by impregnating the ZVI on
the biochar surface. The ZVI particles attached on the biochar surface via
hydrolysis and the reduction method. In this approach, aggressive reduc-
tion by a strong reducing agent as NaBH4 from FeSO4 to FeOHSO4 and
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY 11

finally to FeO imparts magnetic character into biochar. The need for exces-
sive NaBH4 as a prerequisite for the Fe2þ to be completely reduced to FeO,
which has a negative impact on the texture and adsorption characteristics
of the ZVI-MBC. However, the excess amount of NaBH4 causes the dis-
torted surface morphology of ZVI-MBC due to the reduction of the sur-
face, which leads to a reduction in adherence. Hence an optimum molar
ratio (1:10)/FeSO4: NaBH4 could provide compete for reduction and with-
out affecting the textural properties. This may be due to the fact that bio-
char increases the dispersion of the ZVI particles on the porous and coarse
surface of the biochar. This leads to larger available adsorption sites that
lead to an increase. The immobilization of the nZVI on the surface of the
biochar provides a surplus area for adsorption. For achieving better stability
and preventing fair oxidation of nZVI during impregnation, the preferable
approach is doping to biomass after pyrolysis under reducing atmosphere,
which is much preferable compared to the addition before pyrolysis.
To better understand the effect of iron and magnetism in biochar prop-
erties, Jiangfang and coworkers approached the preparation of sewage
sludge-based biochar and test the catalytic efficiency (Yu, Tang, et al.,
2019). The study suggested an increase in porosity and surface area with an
increase in pyrolysis temperatures and simultaneously decreases in oxygen
and nitrogen-based functional groups. The pyrolysis at a higher tempera-
ture also influences the isoelectric point of the biochar in positive direc-
tions (Chen, Kao, Hung, Chien, & Dong, 2018; Huang, Jiang, Huang, &
Yu, 2018; Wang, Liao, et al., 2017; Zhu et al., 2018). The study established
that iron compounds, doped-nitrogen, and graphitic carbon were proved to
be catalytic sites. Similar catalytic studies and reports on different nanoma-
terials and metal oxides were reported, which were impregnated after pyr-
olysis in biomass. For examples, aluminum (Al) modified biochar has been
directly impregnated into straw biochar (Qian, Zhao, & Xu, 2013), manga-
nese oxide loaded biochar (Wang, Sheng, & Qiu, 2015), chitosan loaded
biochar (Zhou et al., 2013), Ni-ZVI-MBC (Devi & Saroha, 2015b), CNT-
biochar nanocomposites (Inyang, Gao, Zimmerman, Zhou, & Cao, 2015)
and ZnS/biochar’s (Yan, Kong, Qu, Li, & Shen, 2015) were reported. In
short, the impregnation of nanomaterials in the biochar before and after
pyrolysis has different effect on the performance and property of biochar.

2.2. Properties of biochar composites


Biochar is known to a versatile material for several environmental remedi-
ation processes (Benedetti, Patuzzi, & Baratieri, 2017). The properties such
as superior porous assembly, high surface area, surface charge, water reten-
tion capacity and a wide range of functional groups (such as –OH, -COOH,
12 S. MANDAL ET AL.

and –NH2 groups) make biochar more operative and holds exclusive import-
ance in the soil remediation as well as improvises the soil properties (Fahmi,
Samsuri, Jol, & Singh, 2018; Li, Khan, et al., 2018). These properties of bio-
char can help in both adsorption and immobilization, which takes place via
complexation, electrostatic interaction, and ion-exchange between contami-
nants (Santos, Brito, Junior, Bonomo, & Veloso, 2019; Xu et al., 2017).
Surface engineering for biochar is implemented based on either significant
requirement or based on the orientation of the application. The common
methodologies that are actively utilized for chemically functionalizing the
biochar are amination, acid/base treatment, impregnation of minerals, surfac-
tant modification, magnetic modification and steam activation (Rajapaksha
et al., 2016). Functionalization of biochar is basically carried out using mate-
rials such as chitosan, carbon-based structures (e.g., graphene, graphene
oxide, and carbon nanotubes), and functional oxide/sulfide nanocrystals
(Bombuwala Dewage, Fowler, Pittman, Mohan, & Mlsna, 2018; Liu, Gao,
Fang, Wang, & Cao, 2016).

2.2.1. Pyrolysis conditions


The pyrolysis conditions have an absolute effect on the properties of the bio-
char. Changes in the biochar properties could be accomplished with a different
temperature, heating level, time of residence and nitrogen flow rate. High-
temperature pyrolysis at 600  C exhibited a high surface area of 108.28 m2/g
with 2.24 nm as pore diameter and also increases the carbon content.
Increasing the pyrolysis temperature increased the pH value and total alkalinity
of the obtained biochar. The concentration of polycyclic aromatic hydrocar-
bons (PAH) varied between 0.16 to 8.73 mg/kg, and the PAH concentration
decreased with an increase of temperature. Prolonging the residence time
increases the biochar surface area, however, there was no correlation found
between the optimized parameters and the PAH content in the biochar (Marıa
et al., 2017). The individual characteristics of the feedstock influence the prop-
erties of the biochar as well. Thus, biochar derived from oak, pine, sugarcane,
and peanut shell at different temperatures varying from 350 to 900  C was
analyzed for its physicochemical properties (Zhao et al., 2018). There was a
significant increase in biochar pH between temperatures 350 to 500  C. The
biochar’s produced varied in three principal components (PC), where, PC1
includes ash, Mn, K, Fe, N, P, and electrical conductivity, PC2 includes O, H,
C, volatile matter, fixed matter, and pH, and finally PC3 includes Ca, Na, and
Mg in biochar. Low pyrolysis temperature influences the PC1 and PC3, which
differs by feedstock type. PC2 was greatly affected by the pyrolysis tempera-
ture. Biochar derived from peanut shell shares both PC1 and PC3 but sugar-
cane was rich in PC1 and low in PC3. Significant PC2 was observed in
biochar derived from wood at an inferior temperature (<500  C) with lower
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY 13

contents of PC1 and PC3. Stable aromatic structures in the produced biochar
are formed at the elevated pyrolysis temperature. This study presents an
understanding of the properties of the biochar derived from different feed-
stock, so the tuning of the surface could be carried out accordingly (Zhang,
Chen, Gray, & Boyd, 2017).

2.2.2. Thermal stability and functional groups


The thermal stability of the biochar is another important parameter that
needs discussion as it governs the sorption performances. The biochar
derived from wood shaving and chicken litter along with activated carbon
was combusted at 375  C for 1 day for the elimination of labile non-
carbonized organic matter. The thermal stability of both wood shaving and
chicken litter was quite poor due to the loss of organic carbon-containing
functional groups. Despite poor thermal stability, higher Cd sorption was
observed due to the negative charge on the surface of combusted biochar,
but sorption capacity decreases for activated carbon which contained 76.5%
pristine C and more negative surface. This could be possibly due to the
minerals presented in the biochars which hold the potential of enhancing
the sorption capacity of heavy metals (Qi et al., 2017). A two year-field
based study was conducted by Liu et al. in red soil region in southern
China (Liu, Xu, Li, & Wang, 2018). The biochar addition did not have a
significant effect over the disturbed soil columns at a high level of water
content. The study also revealed that there is no direct effect over the ther-
mal properties of soil with a change in the soil composition. The biochar
addition exhibited diminishing effects on diffusion and soil’s thermal cap-
acity. The underlying mechanism was understood from the in-situ meas-
urement, which claims that an increase in soil porosity due to biochar
addition is detrimental for the thermal properties of soil, but is beneficial
for increasing in water retention capacity in the soil (Liu, Xu, et al., 2018).
It depends on the soil structure to determine the effect of biochar to be
positive or negative. The physicochemical properties of biochar are also
affecting the grain size and grain yield (Huang et al., 2019). Recently, the
releases of humic substances from the biochar were studied based on the
variance in pH and grain size in biochar. The biochar grain size greater
than 0.5 mm was reportedly having high density and low carbon with a
high content of functional group in the biochar. pH was not influential on
humic acid extraction, rather grain size had a significant effect (Cybulak,
Sokołowska, Boguta, & Tomczyk, 2019).
One of the potential multifunctionality properties of biochar is its char-
acteristics of acting as an effective sorbent for contaminants. A list of bio-
char composites and utilized for various environmental applications were
listed in Table S1.
14 S. MANDAL ET AL.

3. Stability of biochar and its composites in the soil environment


The literature study has obtained sufficient evidence that biochar nanocom-
posites/composites improve the physio-chemical and biological value of the
soil, resulting in an improvement in the agriculture crops yield and remedi-
ation of toxic pollutants efficiency in soil. The following section explained
how biochar and its composites influence the soil microbial community and
soil ecosystem (Doran, 2002; Doran, Sarrantonio, & Liebig, 1996; Doran &
Zeiss, 2000; Laishram, Saxena, Maikhuri, & Rao, 2012). Soil stability refers to
the sensitivity of soil to natural or man-made disturbances. Conversely, soil
flexibility is associated with the ability of soils to restore their quality in
response to any natural or air-induced pathogenic interference (Allen, Singh,
& Dalal, 2011; Lal, 2009). To illustrate the quality of soils and the choice of
sensitive indicators to assess soil quality, it is advisable to select parameters
for management practices. A number of biological properties such as micro-
bial biomass, respiration, amino acids, soil enzymes, and earthworm activity
are considered to be soil quality indicators (Martinez-Salgado, Gutierrez-
Romero, Jannsens, & Ortega-Blu, 2010; Zhou et al., 2018). Overall stability
and volume distribution are guiding principles for assessing the impact of
soil and crop growth practices and dependence on soil quality, as they reflect
resistance to soil erosion. Several studies over the past decade have exten-
sively discussed the obvious effects of the structure, physicochemical proper-
ties and structural relationships of biochar composites in the soil (Table 1).
However, understanding the interaction between biochar and its composites
with soil and microorganisms remain as a research gap.The application of
biochar and biochar based composites to the future soil research is linked in
this review.

3.1. Influence in soil major properties


3.1.1. Influence on soil pH and cation exchange capacity (CEC)
Soil pH and cation exchange capacity (CEC) are inter-dependent parame-
ters, as biochar composites increase the pH-dependent charge in the soil,
which contributes to improved cation exchange capacity (CEC), and by
strengthening the binding to negative parts of the charge function of
organic matter. Thus, in functional areas of organic matter (and biochar),
cation deposition and the formation of the OH-H bonds allow carbon cati-
ons to form weak hydrogen bonds. The properties of biochar composites
vary according to soil aging, especially due to the accumulation of H þ
ions from the soil solutions in the first few weeks and months after soil
amendment supplementation. The extent to which the characteristics of the
biochar changes over time, mostly depend on the source of the biochar and
the type of metal used for the preparation (Gul, Whalen, Thomas,
Table 1. Biochar based composites for application in soil.
Agents/metal
oxides/
Nanomaterials
Biochar composites Biomass and conditions Studies Efficiency/performance Mechanism References
Organic-inorganic Residual bark chips 600  C/2 h 1. Single-Element 1. The adsorption capacity 1. Adsorption Arabyarmohammadi,
composite of Batch Adsorption of MTCB for Cu2þ, Pb2þ, and Khodadadi-Darban,
chitosan, nano- Experiments Zn2þ were much higher et al. (2018)
clay, and 2. Multi-Element than that of the pristine
biochar (MTCB) Batch biochar sample (121.5, 336,
Adsorption and 134.6 mg/g respectively)
Experiment 2. Potential capacity for
3. Immobilization of immobilization of
Cu2þ, Pb2þ, and heavy metals
Zn2þ in Soil 3. Efficient heavy metal sorbent
in mine-impacted acidic
waterand soil
Fe-Mn modified Corn straw KMnO4/Fe(NO3)3 1. Pot experiments 1. Arsenic immobilization 1. FMBCs could Lin et al. (2019)
biochar composite 2. As fractionation 2. increased the soil redox increase the soil
3. Enzyme potential and reduced the redox potential and
activity analysis content of available reduce the content
4. Bacterial of available arsenic
community analysis 2. Amorphous
hydrous
oxide–bound, and
crystalline hydrous
oxide–bound ones
3. Soil enzymes
Clay yak dung Yak dung (YD) 20  C min1 1. Electrochemical 1. Clay enhanced carbon 1. Nanoparticles of Rafiq et al. (2017)
blended biochar /400  C or properties of stability, carboxyl, ketone/ micronutrients and
500  C/2 h. biochar aldehydes, ferrous/ferric the relatively high
clay composites sulfate/thiosulphate, etc concentration of
2. The bulk carbon contents of biochar C/O functional

structure of 2. 500 C showed a positive groups reduce
the biochar impact on yield and quality nutrient leaching
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY

3. Field Trials of bluegrass and 2. N volatilization,


soil properties increase the
abundance of
(continued)
15
16

Table 1. Continued.
Agents/metal
oxides/
Nanomaterials
Biochar composites Biomass and conditions Studies Efficiency/performance Mechanism References
4. Community-level beneficial micro-
physiological organisms and
profile (CLPP) improve
nutrient cycling
S. MANDAL ET AL.

Fe-biochar Rice straw FeCl2/NaClO 1. Porewater 1. significant decrease in grain 1. Biochar-induced Yin, Wang, Peng, Tan,
(Fe-biochar) sampling Cd and As soil As mobilization and Ma (2017)
and analysis 2. biochar exhibited higher was probably
2. Greenhouse pH (10.7) through
experiments 3. Cd in rhizosphere but competitive
elevated soluble As resulting desorption and Fe-
in 468% and 249% biochar-induced
reduction in the root and soil Cd mobilization
grain Cd was probably via
soil acidification
Sulfur and sulfur-iron Bamboo hardwoods 550  C/5 h Pot experiment 1. S-BC and S-Fe BC efficiently – Rajendran et al. (2019)
modified biochar reduced bioavailable Cd
2. Rice growth and chlorophyll
content was increased by S-
BC and S-Fe
BC amendments
3. S-Fe biochar is an effective
way to prevent Cd
accumulation in rice grains
MgO-coated corncob Corncob 600  C 1. Lead 1. CB 0.07 (m2/g) 1. Surface complexation Shen et al. (2019)
immobilization 2. 82.95 ± 0.17 (%)
2
in soil 3. MCB 26.56 (m /g)
2. Toxicity 4. MCB 53.51 ± 0.25 (%)
Characteristic
Leaching Procedure
Biochar and 1. Eucalyptus. 2. 500  C/2 h 1. Greenhouse 1. Biochars led to increased soil 1. Surface complexation Penido et al. (2019)
sewage sludge Sewage sludge experiment and leachate pH
2. Biochars reduced Cd, Pb,
and Zn bioavailability
Manganese oxide- Corn straw 600  C/2 h 1. Pot experiment 1. Mn oxide-modified 1. Adsorption Yu et al. (2017)
modified biochar KMnO4 2. Arsenic biochar application reduces
concentrations and arsenic concentration in
species and amino brown rice
acid content 2. Mn oxide-modified
3. Extraction of soil biochar composites have a
arsenic and iron- positive effect on amino
manganese plaque acids and improve
from roots rice growth
3. Mn oxide-modified biochar
application can be used to
remediate arsenic-
contaminated soil
Ferro-manganese Corn straw 600  C/2 h 1. Pot experiment 1. FMBC improve the biomass 1. Adsorption Lin et al. (2017)
oxide impregnated KMnO4 2. Speciation of As in weight in As-
biochar rice and soil contaminated soils.
composites 3. Extraction of Fe 2. The addition BC
and Mn plaques decreased the total As
from roots concentration
3. applicability as reducing
the bioavailability of As in
As-contaminated soils.
4. Addition of FMBC
increased the ratio of
essential amino acids in
the grain
Biomass-derived Palm fiber 600  C/N2 1. Pot incubations 1. Biomass-derived magnetic 1. Electrostatic Cui, Jin, Li, and Li (2019)
magnetic 2. Removal of As(III) nanocomposite (BMN) was attractions
nanocomposite from fabricated through one- 2. Oxidation
(BMN) soil suspension step pyrolysis 3. Immobilization
3. Removal of As(III) 2. BMN could be supported by
from sandy soil sponges and filter papers,
4. Leaching behavior which could efficiently
of As(III) remove the As(III) from soil
Biochar/clay/chitosan Clay/chitosan 600  C/2 h 1. Soil 1. Studies on Pb-Zn – Arabyarmohammadi,
nanocomposite acetic acid sequential mine soils Darban, van der Zee,
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY

(MTCB) extraction 2. MTCB contribute to OM and Abdollahy, and


2. Column leaching increases the Ayati (2018)
exchangeable fraction
17

(continued)
18

Table 1. Continued.
Agents/metal
oxides/
Nanomaterials
Biochar composites Biomass and conditions Studies Efficiency/performance Mechanism References
Biochar and lime Rice Straw/Lime 500  C/2 h 1. Sequential 1. The soil pH is increased 1. A higher Li, Xu, et al. (2018)
extraction of with biochar addition concentration of
soil Pb 2. RBL decreased CaCl2- soluble SO42
2. Rice harvest extractable soil Pb more decreased Pb
S. MANDAL ET AL.

and analysis effectively than biochar availability in


3. Distribution 3. RBL decreased Pb the soil
experiments of Pb, content in brown rice 2. Immobilization
Fe, Mn, Zn, Cu, and
Se in rice grains
Magnesium Commercial product 600  C/1 h Phosphorus leaching 1. Phosphate-P losses from the Adsorption Riddle et al. (2018)
(hydr)oxide- Ecoera agricultural four control treatments were
coated biochar crop residues in the order of
organic > organic >
sand > loam
2. organic soils have a greater
potential than mineral soils
for releasing loads of P to
surface waters
Sludge biochar/ Willow 700  C Co-application of biochar Biochar can uphold the transition Binding Bogusz and
sewage sludge (Salix viminalis) with sewage sludge of available forms to their Oleszczuk (2018)
for speciation of Ni residual forms
and Zn
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY 19

Sachdeva, & Deng, 2015; Lehmann et al., 2011; Zhu et al., 2017). An
increase in soil pH by adding biochar and biochar composites reported for
across many soil types (Stewart, Zheng, Botte, & Cotrufo, 2013; Tang et al.,
2019; Xu, Sun, Shao, & Chang, 2014). Other research shows that high pH
biochar increases the soil pH by about one-third of lime and increases the
calcium levels and reduces aluminum toxicity on red iron soils (Agegnehu,
Srivastava, & Bird, 2017). Fewer reasons are reported, such as negatively
charged phenolic, hydroxyl and carboxyl groups on the biochar surfaces
that bind Hþ from the soil solution (Chintala, Mollinedo, Schumacher,
Malo, & Julson, 2014; Lu, Gong, Tang, Huang, & Gao, 2015; Sheng, Zhan,
& Zhu, 2016). Wang et al. reported on the development of an acidic soil
recovery agent (ASRA) based on a nanocomposite consisting of anhydrous
sodium carbonate (ASC), attapulgite (ATP), straw ash-based biochar and
bio-silica (BCS) (Wang, Guo, et al., 2017). Perhaps it is the loss of Ca2þ
and the increase of acidic pH in soil, because biochar nanocomposite acts
as an inhibitor with a porous nano networks. However, the pH effect may
also depend on the content of the SOM, which is the main determinant of
the soil CEC (Yu, Zhou, Huang, Song, & Qiu, 2015).
Compared with the effect of biochar on the improvement of the pH value
of acidic soil, the pH value of biochar treatment of alkaline soil improved
slightly (Gonzaga et al., 2018). In addition, the effect of biochar on soil pH
and CEC values depends on the type of soil. Gai and Zhai and other studies
found that the addition of corn stover biochar can significantly increase the
red soil pH, but the effect on black soil pH is not significant, and the pH of
the red soil increases as the amount of biochar application increases (Gai &
Zhai, 2017). The application of biochar and biochar based composites is bene-
ficial to reduce soil bulk weight, improve soil pH, organic matter content,
NO32 content, effective phosphorus content and water content, and showed
that the application of biochar and its composites as soil improver to farmland
can effectively improve soil physicochemical properties and improve cultiva-
tion performance. The extent to which biochar properties change over time
depends on biochar sources and soil types, as well as climatic conditions,
which are still need to be studied (Heitk€otter & Marschner, 2015).

3.1.2. Influence on soil bulk density, porosity, water retention capacity, and
redox conditions
Soil bulk density, porosity, water retention capacity, and redox conditions
have a significant influence on different interdependent properties of the
soil, including microbial abundance and plant growth (Krishnakumar et al.,
2014). A.M Goodman and A.R Ennos have reported that soil with high vol-
ume density (>1.6 mg cm3) has a low capacity to absorb water and has
great penetration resistance to plant roots eventually entering the soil, and
20 S. MANDAL ET AL.

plant growth will be affected (Goodman & Ennos, 2007). Githinji et al.
determined that the bulk density was also significantly reduced by increas-
ing the use of biochar (Githinji, 2014). In a report by Berihun and others,
the application of biochar had a significant effect on soil bulk density and
porosity. The untreated soil has a large volume density (qb ¼ 0.91 ± 0.01%,
low porosity ¼ 0.67 ± 0.01%) than 18 t ha1 of Lantana biochar (qb¼
0.72 ± 0.02% and porosity ¼ 0.74 ± 0.00%). However, this change in poros-
ity is negligible, and the effect is due to a decrease in the bulk density of
biochar modified soils, but increased the soil porosity and soil aeration
(Berihun, Tadele, & Kebede, 2017).
According to Mukherjee et al., the application of biochar reduces the
bulk weight of the soil because of the very high porosity of biochar, which
significantly reduces the bulk density of the soil by increasing the pore vol-
ume (Mukherjee, Lal, & Zimmerman, 2014). The pore space or porosity of
soil refers to the pore volume ratio of the total soil volume. The application
of biochar increases the net porosity of the soil, but the increase in porosity
depends on the type of biochar used and the type of soil, Furthermore, the
increase in total porosity is usually attributed to the low density and highly
porous inner structure of biochar particles (Herath, Camps-Arbestain, &
Hedley, 2013). Hence, it remains very important to understand the nature
of biochar and its composites application in soil. Similarly, Jian and cow-
orkers reported that the bulk density of the soil changed by biochar
amendment which was observed to be 1.10 mg/m3 during soil incubation
(Jien & Wang, 2013). However, within 21 days of incubation, the bulk
density increased rapidly in control, then the constant value of about
1.40 mg/m3 was maintained, and the bulk density of the biochar modified
soil decreased significantly from 1.42 mg/m3 to <1.15 mg/m3 until 105 days
after incubation, and the rate of reduction increases with the increase of
the application rate of biochar. In addition, the application of biochar in
the soil can be significantly increased, and the soil aggregation rate of the
soil modified has been higher than that of the control soil. However, there
were significant differences observed between the altered soil and the con-
trol soil after 84 days of incubation and all treated soils have found an obvi-
ous peak, occurring in the 21 days. This clearly reports the change in soil
property after the incorporation of biochar and its composites. Indifference
to this, the report is also suggested in applying biochar and planting sys-
tems. The reason for the determination is that the bulk density of the bio-
char modified soil decreases and this may have a positive effect on root
and microbial respiration.
It also has been reported that, over time, biochar addition to the soil
could increase water and nutrient content, as the surface oxidized over
application time and biochar properties seem to be hydrophobic in nature,
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY 21

which further causing higher water absorption and retention capacity. The
application of biochar has efficiently changed the water holding capacity of
soil by changing soil porosity and agglomeration level. Fang et al. report
suggested that compared to untreated soil, the biochar amendment at a
rate of 0 thm2 and 100 thm2 significantly reduced soil bulk density by
14.6% and 32.5%, respectively (Fang, Li, & Bin, 2014). The decrease in soil
bulk density indicates an increase in soil total porosity, which means the
improvement of soil aeration and the increase of soil moisture permeability
(Jeffery, Verheijen, van der Velde, & Bastos, 2011). Uzoma et al. investi-
gated that, the application of biochar increases the available water content
of the soil to 97% and the saturated water content up to 56%, but the
increase depend on the surface of biochar (Uzoma, Inoue, Andry, Zahoor,
& Nishihara, 2011). Besides, it is ascertained that there were hydrophilic
functional groups present on the surface of the graphene sheet of the bio-
char and also on the pores and the significant increase in soil water reten-
tion capacity by biochar application in case of sandy soil is higher than
loamy and clay soil. Also, the water retention capacity and soil profile are
considered to depend on soil depth, gravity drainage, and evapotranspir-
ation. Soil water retention curve (SWRC) indicates the relationship between
soil water pressure (pressure head) and volume water content (Geroy
et al., 2011).
The large specific surface area and a hydrophilic group of biochar make
its hygroscopic capacity 1–2 orders of degree higher than that of soil
organic matter and biochar can effectively reduce the tensile strength of the
soil and reduce cracks (Accardi-Dey & Gschwend, 2002). The settlement
caused by soil contraction, thus improving the water holding capacity of
the soil (Meng et al., 2017). In compare to the above mentioned fundamen-
tal properties, soil redox condition also has been influenced by the addition
of biochar composites. In a study by Lin et al., redox properties were
affected when using biochar (BC) and Fe-MN modified biochar (FMBC) to
treat contaminated paddy soil. Compared with the control, the redox
potential of the samples treated with BC and FMBC increased suggestively,
and the above potential changes were positively interrelated with the pro-
portion of preparation materials (BC and FMBC). Iron and manganese
have a great effect on the redox process in soil, and even affect the behav-
ior of soil microorganisms (Lin, Li, Liu, Qiu, & Song, 2019). Other
researchers have recently reported similar observations (Burrell, Zehetner,
Rampazzo, Wimmer, & Soja, 2016; Huang et al., 2016; Liu et al., 2013; Sun
& Lu, 2014).
Hence, it is clearly agreed with the above statements that, soil porosity,
water retention capacity, soil redox conditions and bulk densities are inter-
dependent parameters and influence to each other to a very high extent.
22 S. MANDAL ET AL.

3.1.3. Effect on soil organic matter (SOM)


Soil organic matter (SOM) is the primary source of soil nutrients. SOM
can promote the activity of soil microbes and the absorption of nitrogen
and reduces nitrogen losses via volatilization (Mary, Recous, Darwis, &
Robin, 1996). Dynamic soil organic matter (DSOM) is an important com-
ponent of soil that affects dissolved organic carbon (DOC), microbial bio-
mass and light-fraction organic matter. DSOM is also related to soil CEC
and pH values affected by soil texture, moisture content, temperature, and
other factors (Nannipieri et al., 2017; Reeves, 1997; Tan et al., 2017).
Biochar and biochar composites itself is enriched in organic carbon,
important metals and aromatic. The addition of biochar to the soil, could
somewhat improve soil humus and organic content, as well as indirectly
increase the proportion of soil C/N ratio by absorbing other nutrients
(Munda et al., 2018; Sandhu et al., 2017).
Fang et al. (2014) considered that biochar and its composites can improve
the stability of the agglomerates and reduce the leaching of organic matter
by promoting the formation of soil organic-mineral complexes, and micro-
bial growth, wrapped or adsorbed in the biochar voids and organic-mineral
complexes. The amount of biochar carbon applied and the type of soil will
affect the amplitude of its change to soil organic matter. Increasing the appli-
cation rate of biochar and its composites increases the carbon dioxide of soil
respiration, and the carbon dioxide released by the soil decreases over time
(Elzobair, Stromberger, Ippolito, & Lentz, 2016). The pH value is an import-
ant factor in controlling soil CO2 emissions. In acidic soils, the effect of bio-
char on CO2 emissions is more significant. Wu, Senbayram, et al. (2018)
studied that the application of corn-based biochar had no significant effect
on the CO2 emissions of alkaline clay, but the CO2 emissions of acidic sandy
soil were reduced by 26.9% and it indicated that in acidic soils, biochar is
more likely to affect CO2 emissions. Biochar and its composites also increase
the content of total nitrogen in the soil, which is mainly caused by the
release of nitrogen from biochar into the soil (Gai & Zhai, 2017). The poten-
tial for nitrous oxide reduction of biochar in acidic soils depends on the
presence of soil NO3 (Wu, Senbayram, et al., 2018). Schomberg H et al.
studies have shown that increased soil pH under the application of biochar
could promote ammonia evaporation (Schomberg et al., 2012). In addition,
Nelissen V et al. found that biochar/biochar composites can promote the
mineralization and nitrification of soil nitrogen, which shows that the con-
tent change of NH4þ-N is resulting from the accumulation of various effects
(Nelissen et al., 2012). Moreover, different biochar and biochar based com-
posites have different effects on nitrogen conversion, and the effect of straw
biochar on nitrogen conversion rate is greater than that of wood-based bio-
char (Li, Liu, Li, Jiang, & Wu, 2016).
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY 23

The application of biochar significantly increases the average total nitrogen


content of the soil. In addition, the biochar impregnated calcium alginate
(CA) bead complexes are introduced, which has the characteristics of
improving moisture and nutrient retention and good control release. CA and
calcium alginate ball-milled biochar (CA-BMB) (CA-BMB) released more
than 80% and 90% of the Kþ content after 24 hours. Compared to CA beads,
CA-BMB beads release less K þ , indicating that the impregnation of CA
with BMP increases the retention of Kþ. The reason for the release is mainly
controlled by the diffusion in the pore networks of beads in CA and CA-
BMB. Compared to CA beads, the CA-BMB beads show less NO3 release,
indicated that the CA-BMB could retain more NO3 after impregnating with
BMB (Wang, Gao, Zimmerman, Zheng, & Lyu, 2018). In addition, combin-
ing the biochar composites with available nitrogen sources can create a sus-
tainable material to improve soil properties and fertilization. The use of the
biochar-ash composites as fertilizer brings an instant financial incentive and
improves environmental sustainability, while long-term positive effects are
expected from soil advances of biochar (Berihun et al., 2017; Li, Hu,
et al., 2018).
The existing literature seems to indicate that the C supply in the soil, fol-
lowed by pH, porosity, CEC, and aeration, can be increased or decreased by
adding biochar. A conceptual outline is proposed for illuminating the inter-
action between plants and biochar/biochar based composites in the soil is
presented in Figure 4. In the soil of concern, the leaching or leaching of
nutrients can indirectly improve the efficiency of nutrient use by adsorption
of essential elements. Biochar can also directly controls the acquisition of
plant root nutrients by changing soil nutrient content as a source of
nutrients. However, there is a lack of research on biochar composites and
metal-assisted biochar, the effects on soil nutrient retention and properties.
3.2. Influence in soil microbial and enzyme activity
The physical and chemical properties of biochar and biochar composites, as
well as soil properties changes caused by biochar addition, can amend the
soil micro-organisms and enzyme activities. It is studied that biochar pores
provide habitat for micro-organisms and simultaneously improve the dens-
ity, pH, and help in the flow of air, water, and nutrients within the soil
environment. Li et al. and Ling et al. observed and reported that soil prop-
erties such as pH and organic matter, available phosphorus, iron, manga-
nese, and other oxide content were widely variable (P < .05), thus being
responsible for the variation of the bacterial community in these soils (Li,
Zhang, et al., 2016; Ling et al., 2016). The chemical properties of biochar
and its composites, which can influence microbial growth within the bio-
char pores, create an apparent load, alter the nutrient concentration and
24 S. MANDAL ET AL.

Figure 4. Conceptual outline of the interaction between plants and biochar/biochar based com-
posites added to soil influencing soil terrestrial properties (arrows are meant to indicate effects
and flow).

dynamic organic carbon that desorbed or solubilized from biochar.


Uchimiya, Ohno, and He (2013) report that biochar extract derived from
various feedstock sources, having fulvic and humic content, which is simi-
lar to soil organic carbon and found to be thermally stable. This research
gap presents a distinctive opportunity to examine the modification of bio-
char surface and elucidate the impact on the soil microbial community.
Similarly, soil enzymes mainly come from soil microbial secretion, plant
root secretion, and residual decomposition, and participate in all biochem-
ical reactions in the soil, its activity reflects the activity of soil micro-
organisms and the ability of soil nutrient transformation, which maintains
the soil health and provide plants with the nutrients needed for growth
(Wu, Senbayram, et al., 2018). Application of the biochar and its compo-
sites may improve the soil enzyme activity, which further depend upon the
soil environmental conditions. In most of the cases the soil enzyme activity
increased with the addition of biochar and its composites (Masto et al.,
2013). The activity of the urease and catalase can be significantly improved
by the bamboo, straw, and straw-based biochar, and there was no signifi-
cant change in the phosphatase activity (He, Zhong, & Yang, 2017).
Since biochar has strong adsorption capacity, its effect on soil enzymes is
more complex. On the one hand, the biochar can adsorb the reaction sub-
strate, promote the enzymatic reaction and improve the soil enzyme activ-
ity. On the other hand, biochar can adsorb enzyme molecules, the
enzymatic reaction binding sites to protect, thus averting an enzymatic
reaction, and then reduce soil enzyme activity (Lehmann et al., 2011). Jien
and others suggested the changes in soil microbial activity after the applica-
tion of biochar, and microbial biomass carbon (MBC) content at 0 d, 21 d,
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY 25

63 d, and 105 d of incubation. The studies have shown that the use of bio-
char greatly increased MBC at the beginning of incubation, for 63 d and
105 d (only in 5% application rates). The difference was statistically import-
ant (p < .05), with the exception of the results of the analysis of the 21 d.
Additionally, each treated soil had the highest levels of MBC, the highest
MBC content per soil, 5% of biochar modified soil (3200 mg kg1), 2.5% of
biochar modified soil (1145 mg kg1) and the controlled soil (1759 mg
kg1) (Jien & Wang, 2013). Similar studies are well-documented suggesting
influence in microbial and enzyme activity (Foster, Hansen, Wallenstein, &
Cotrufo, 2016; Jien & Wang, 2013; Lin et al., 2019; Zheng et al., 2016).
Furthermore, these studies show that the nutrients in biochar and bio-
char composites, including nutrients from biochar and nutrients from
bulk soils to biochar substrates, control the growth and activity of micro-
organisms, as well as the amend/obstruct enzyme secretions. The changes
in the soil microbial community structure and its activity in biochar and
its composites should be considered as the physiological and biochem-
ical factors.

4. Remediation of heavy metals in soil


Soil pollution is hazardous as it is the base of human survival leading
through the food chain. It is the versatility in the soil that makes it more
prone to polluted by heavy metals. However, the development of the car-
bon-based biochar materials with functionalization has emerged as a poten-
tial medium to stabilize and remediate the soil-pollution caused by these
hazards that land in our vicinity. As already mentioned in the previous sec-
tions, there are certain advantages of the nanocomposites of biochar over
the pure biochar, which includes increased adsorption sites (Devi & Saroha,
2015a), increase in reactivity with soil contaminants (Yang et al., 2016),
enhanced thermal stability, and also less agglomeration could be achieved
with the incorporation of magnetic nanoparticles with the higher surface
area for reaction (Ye, Zeng, et al., 2019). It has been reported by several
researchers that loading of nanometals onto the surface and into the pores
of biochar changes the properties of the resultant biochar with the
increased porosity, surface charge, and surface functional groups (Ye, Yan,
et al., 2019). Furthermore, it also depends on the type of nanometals on
the surface of biochar for complexation with other pollutants in the soil
such as heavy metals. The consideration of valence state, electronegativity,
redox potential and ionic radii of the hydrated complex is also important
for understanding the efficiency of the interaction (Ho, Zhu, et al., 2017;
Yao, Gao, et al., 2011a, 2011b).
26 S. MANDAL ET AL.

4.1. Arsenic (As)


As environmental impacts are attracting worldwide attention due to arsenic-
induced groundwater and soil contamination. Arsenite and arsenate are the
dominant forms of As in soil. Studies were conducted to remove arsenic in
order to avoid the involvement of arsenic species in the food chain (Mitra,
Chatterjee, Moogouei, & Gupta, 2017). In this direction, biochar is modified
with iron and studied for immobilization, considering that iron has good
affinity toward arsenic species (Wang, Gao, et al., 2015). Ch et al. investi-
gated the remediation studies of arsenic-contaminated paddy soil by biochar
composites. Biochar is improvised in terms of surface properties through Fe-
oxyhydroxide sulfate (Biochar-FeOS), FeCl3 (Biochar-FeCl3), and zero-valent
iron (Biochar-Fe). These materials were utilized for incubation experiments,
where soil is highly contaminated by arsenic (122.4 mg/Kg). The three
materials had minimal effect on the soil pH. As per the sequential extraction
analysis data of soils with the incubation time from 7 days to 90 days, the
modified biochar produce immobilization effect as compared to pure bio-
char. However with the increase in the incubation time, the immobilization
effects of the three types of iron-modified biochar (BC-FeOS, BC-FeCl3, and
BC-Fe) were significant, and the fixation rate reached 13.95 -30.35%, 10.97 -
28.39%, and 17.98 - 35.18%, respectively. The respective decrease at longer
period of incubation time is reasoned for soil redox conditions. The oxida-
tion of iron is also a common factor that decreases the affinity at later stage
of immobilization which desorbed the adsorbed species (Wu, Cui,
et al., 2018).
In a similar study by Lin et al., focused on Fe-Mn modified biochar com-
posite (FMBC: FMBC1 and FMBC2) treatment on the pH, redox properties,
enzyme activities, and bacterial communities of As-polluted paddy soil. The
effect was significantly observed and the two utilized modified biochar
composites affect soil pH, and increased the soil redox potential and
reduced the available content of arsenic in the soil. Furthermore, the modi-
fied biochar increased the soil enzyme activities, with the exception in
FMBC2. The FMBC2 application decreased alkaline phosphatase (ALP). The
pH influences the amount of negative charge on the soil surface and bio-
availability of AsO43 or AsO33. In addition, soil organic matter can form
heavy metal complexes in order to encourage attachment. Results obtained
show that various arsenic forms can be converted to stable form under
suitable conditions. The treatment of BC or FMBCs resulted in the help
and relocation of arsenic in the soil and nature (physical and chemical
properties) of soil has a great influence on the distribution of As species
(Lin et al., 2019). Similar research was well reported for arsenic remedi-
ation in the soil (Kim et al., 2018; Qiao et al., 2018; Rocco et al., 2018).
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY 27

4.2. Cadmium (Cd)


Cd is a nephrotoxic heavy metal and has no physiological and biochemical
benefit for plants, yet easily accumulate in crops, leading to a number of
diseases in livestock and humans. The immobilization of cadmium from
soil could reduce the bioavailability. Biochar and its composites provide
suitable properties for in-situ remediation. In a two-year field trial in paddy
soil, it was observed that under flooded conditions, the mobility of cad-
mium decreases and application of zero-valent modified biochar with silica
sol successfully reduced the mobility of cadmium species in the soil. The
grain yield was also observed and a decreasing trend was observed in order,
iron-modified biochar plus silica sol (Fe-BC plus Si) > silica sol (Si) > Fe-
BC > control (CK). The studies suggested that the addition of iron modi-
fied biochar had a good stabilization capacity of Cd and significantly
increased the soil pH in a two-year duration from 4.6 to 5.1. The reason
was that the biochar enhanced the complexation and adsorption of metal
cations on the surface of the biochar and reduced metal mobility by
increasing the pH of the acid soil. In addition, foliar application of the sili-
con deposition on the cell walls limited the transport of Cd onto the
plasma membrane by forming (silicon-hemicellulose matrix)-Cd complex
and thus inhibited Cd uptake. However, this study well unified the modi-
fied biochar with silica sol and reduced the available Cd content in the soil,
but unsuccessful to discuss the individual influence effectively (Pan, Liu,
Yu, & Li, 2019). In a similar report, studies reported remediation of cad-
mium in soil by biochar-supported iron phosphate nanoparticles. Dodium
carboxymethyl cellulose (CMC@BC@ Fe3(PO4)2) composite was used for
the study and bio-accessibility to cadmium was reduced by 80% and the
transition from easier extractable Cd to the less available form was based
on a physiological extraction test. It was observed that at the short span
reaction time (1 day), the efficiency of the composite material was higher
than the sum of the individual remediation efficiencies, possibly because Fe
was adsorbed on the surface of biochar in the composite material and
occupied the Cd adsorption sites. Furthermore, it was reasoned that the
oxygen-containing functional groups on the surface of the biochar formed
complex with Cd to increase the organic matter fraction,which was consid-
ered as the main mechanisms to remediate heavy metal in soil.
Additionally, the composite has also been applied to the soil and not only
decreased Cd bioaccumulation of the cabbage mustard plants, but also
allowed plants to develop and grow (Qiao et al., 2017). Similarly, other
reports are also well documented and suggested that surface functional
properties and pH conditions are governing functions in cadmium immo-
bilization in the soil (Gao, Peng, Zhou, Adeel, & Chen, 2019; Tariq,
28 S. MANDAL ET AL.

Samsuri, Karam, Aris, & Jamilu, 2019; Wang, Gu, et al., 2019; Xu, Fang, &
Tsang, 2016; Yao et al., 2019).

4.3. Lead (Pb)


Pb is considered to be one of the most toxic elements and a potential
threat for both humans and animals, because of extensive manufacturing in
paint and mining sectors, lead pollution is often recorded in soils around
the world (Ahmad et al., 2017). Immobilization measures are therefore
essential for the protection of human and animal health in contaminated
soils. Biochar and its composite present effective in-situ remediation for
lead in soil. The lead immobilization mainly influenced by soil pH, and
extensive measures were taken to increase the pH of the soil to alkaline
nature and promising results were obtained (Zhang, Wang, et al., 2013).
The mechanisms for immobilizing Pb for each biochar also differ and
depend mainly on properties, such as pH, ash content, aromaticity, surface
area and prominent surface functional groups (e.g., -C ¼ O, C ¼ C, -CH,
-OH, -COOH) (Igalavithana et al., 2017). Igalavithana and co-researchers
studied the short-term incubation studies for soil immobilization by using
a set of eleven biochar (vegetable wastes, pine cone, and crop residues).
This study presented a useful insight on the behaviors in terms of reactivity
and stability of biochar toward heavy metals in soil. The short term study
was considered to identify the potential improvements for long term
appraisal. The study suggested that nitrogen-containing functional surface
groups, particularly amine (-NH2), could have facilitated Pb immobilization
by strong covalent bonding and an increased number of cation exchange-
able sites by the presence of amino groups on the biochar surface.
Furthermore, it was also suggested that considering the feedstock from red
pepper stalk, vegetable waste, vegetable waste þ pine cone (1:1) and wood
bark biochar’s produced at high temperatures in comparison to low tem-
peratures showed higher Pb immobilization (Igalavithana et al., 2019).
Similarly, different feedstocks including soybean stover (SS), peanut shells
(PS) and pine needles (PN) based biochar were tested for lead immobiliza-
tion. It was observed that increase of biochar addition significantly increase
the soil pH and Ec in both shooting and agricultural soils. This could be
reasoned for increase in ionic activity in soil because of cations (Ca, Mg,
K) and anions (Cl, PO4, and SO4) released by the biochar. Higher oxygen
content and biochar produced at 300  C presented effectiveness in immobil-
ization toward Pb and other divalent metals. The high effectiveness is
mainly because of multiple mechanisms including surface complexation,
and precipitation (Ahmad et al., 2017). In another study where a chitosan/
clay/biochar nano-biocomposite (MTCB) was chosen to immobilize lead in
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY 29

acidic soil, with having high surface area and mesoporous structure in
compare to pristine biochar, it facilitates Pb immobilization as binding
with surface-active NH2 groups (Arabyarmohammadi, Khodadadi-Darban,
et al., 2018). Similar studies were well discussed for lead immobilization
(Cao, Ma, Liang, Gao, & Harris, 2011; Houben, Evrard, & Sonnet, 2013;
Mahar, Wang, Li, & Zhang, 2015; Moon et al., 2013; Rizwan et al., 2016;
Rodriguez, Lemos, Trujillo, Amaya, & Ramos, 2019; Shen et al., 2019).
Overall, all the mentioned studies presented a suitable opportunity to
immobilize Pb from Pb polluted soil, however the simultaneous application
of biochar and its composites in soil could induce positive or negative
effects toward soil microbial community is missing to higher extent.

4.4. Mercury (Hg)


Hg is a hazardous heavy metal rated to be one of the “ten leading risk
chemicals” and is of significant concern worldwide owing to its toxic effect
on human health. Hg can be transported and thoroughly mixed in its elem-
ental form (Hg(0)) when released into the atmosphere. The main sink of
elemental mercury is the deposition of divalent mercury (Hg(II)) to soil or
water bodies after oxidation (Wang et al., 2020). Immobilization techniques
are promising stabilization techniques, which prevent mercury migration
by chemical complexation. Becker’s and coauthor studied the impact of
biochar (two different pine cone biochars) on immobilization of total Hg,
methyl-mercury (Me-Hg), and ethyl-mercury (Et-Hg), under redox condi-
tions. The research makes unsatisfactory conclusions regarding biochar
effects, but the resulting observation is crucial to obtaining standard infor-
mation in future studies. The report suggested that there is no decrease in
concentrations of total Hg, MeHg, and EtHg after soil treatment with the
various biochar, apparently because possible binding sites for Hg are occu-
pied by other ions and/or blocked by the biofilm (Beckers et al., 2019).
Moreover, there are also positive reports where biochar surface is enhanced
and remediation of mercury was carried out. Sulfur-modified rice husk
(RH) biochar is being used and sulfur modification onto biochar surface
increased Hg2þ adsorptive capacity by  73% to a very high Hg contami-
nated soil (1000 mg/Kg). It was observed that sulfur inclusion in biochar
surface increases the ash content by 43.43 (% dry wt.), pH, 10.65 and dras-
tically reduced the pore volume by 12%. The reason for incorporation of
S is mainly strong binding affinity toward sulfur and forming HgS stable
phase. pH driven the immobilization efficiency, a lower pH is found to be
suitable for high Hg immobilization, moreover S-modified biochar was
more effective than pristine RH biochar. However, the study didn’t con-
sider the microbial impact on sulfur consuming bacteria present in the
30 S. MANDAL ET AL.

contaminated soil (O’Connor et al., 2018). Similar, studies are well reported
for Hg immobilization in soil (Lyu, Gong, Tang, Huang, & Wang, 2016;
Yin, He, Zeng, & Chen, 2016). However, a detailed study needs to be
offered in near future considering the soil environmental factors in terms
of microbial influence for Hg immobilization.

4.5. Chromium (Cr)


The presence of heavy metals in soils presents a significant threat to the
environment, especially chromium, because it is the most common indus-
trial commodity element and highly utilized in electroplating, wood preser-
vation and leather tanning. Cr(VI) and Cr(III) are the two oxidation states
of Cr, because of its carcinogenicity, persistence, and bioaccumulation, Cr
contamination causes the growth of Cr in plants, enters the food chain
through plants and poses potential hazards to human health. Thus, it is
important to reduce the potential toxicity of Cr(VI) in soil. Biochar and its
composites present significant opportunities for the purpose. In this direc-
tion, Lyu et al. had made a successful effort in immobilizing Cr in soil by
using nanoscale iron sulfide supported biochar composite. It was under-
stood that surface sorption and reduction/precipitation are dominant
immobilization mechanisms and easily convert most accessible Cr into less
accessible forms. The toxicity characteristic leaching procedure (TCLP)-
leachable Cr(VI) concentrations in biochar treated sample showed signifi-
cant changes. The CMC-FeS@biochar was more successful in immobilizing
soil Cr(VI) than bare FeS, which was due to the larger specific surface area
and also offered higher capacity for Cr(VI) immobilization than plain bio-
char, while biochar had a larger specific surface area than CMC-
FeS@biochar, indicating the significant role of FeS over 180 days. Surface
complexation between the Cr(VI) and COOH functional groups as well as
the formation of CO by the chemical redox reaction between the Cr(VI)
and the C bond on the CMC-FeS@biochar surface. The increased oxygen-
containing functional groups, derived from biochar and CMC, may
improvise Cr(VI) immobilization. However, the study also suggested
enhancement to soil organic matter and microbial activity by the addition
of CMC-FeS@biochar, but there is a very strong lack of information about
sulfur and its chromium reactivity on the soil (Lyu et al., 2018). The indi-
vidual roles of every component in biochar have a strong link to the
enhancement for future applications of composites based on biochar. A
similar trend of study reported by Su et al. also used nanoscale zero-valent
iron supported biochar, where nZVI acted a reactive component and bio-
char provided a suitable environment. It was observed that nZVI@BC
exhibited better stability and mobility than that of bare-nZVI. Furthermore,
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY 31

nZVI@BC reduced the leachability and bioavailability of Cr in the soil. The


immobilization effect of nZVI@BC on total Cr in the soil was slightly
higher than that of bare-nZVI because negative charges or functional
groups on the BC surface could prevent Cr(III) easily. According to physio-
logical based extraction test (PBET), the bio-accessibility of the total Cr
was further declined when treated with nZVI@BC in compare to only bio-
char. Additionally, nZVI supported biochar also benefited to the plant
effectively promote the seedling growth (pot experiments). This study pre-
sented notable observations and offer possible opportunities to utilize for
large scale applications (Su, Fang, Tsang, Fang, & Zhao, 2016). In the same
way, other reports are well documented and present a strong possibility for
the use of biochar and its composites for Cr removal in the soil (Alaboudi,
Ahmed, & Brodie, 2019; Gomes, Dias-Ferreira, & Ribeiro, 2012; Zhang,
Fang, & Zhang, 2017; Zhang, Zhang, & Fang, 2018). However, there are
major difficulties in transformation from lab scale to large scale treatments,
such as the reactivity of iron nanoparticles and the stability toward micro-
bial communities and soil animals. It is, therefore, necessary to present a
scientific analysis considering the effects in the microbial community.

5. Immobilization mechanism
Multiple mechanisms have driven the efficiency of immobilization and
elimination of heavy metals and toxic pollutants in the soil. Biochar is
widely reported to be highly efficient in soil restoration, since inorganic
pollutants are highly adsorbed and the biochar often restores nutrients,
controls the bacterial activity, and soil biota is maintained. Nevertheless, lit-
tle is known of the use for the same reason as biochar-based nanocompo-
sites. A detailed mechanistic process for the removal of toxic heavy metal
ions by utilizing biochar composite is presented in Figure 5.

5.1. Adsorption
The abundance in the surface functional groups in biochar is prone to
physical/chemical adsorption forming metal-ligand complexes due to their
affinity toward heavy metal ions or chemical reactions. Physical adsorption
is used to adsorb metal ions in micro pores with surface characteristics
(surface area and porosity), and it is closely related to the pollutant levels
and temperatures of the system. Additionally, the surface containing oxides
of iron or hydroxides can form coordination bonds and the heavy metals
get adsorbed onto the functional group as Fe-O-Mnþ (Ma et al., 2015). In
such cases, the surface functional groups such as Fe-O, or Fe-OH are the
main contributing mechanisms governing the removal of heavy metal ions
32 S. MANDAL ET AL.

Figure 5. The mechanism involved and properties of biochar-based composites employed in


the removal of toxic pollutants in the soil.

(For, ex., Zn2þ, Pb2þ, Cu2þ) (Han et al., 2015). The functional groups can
also undergo a change in the properties due to changes in the pH affecting
the uptake of heavy metal ions. Thus, the deprotonation of the functional
groups on the biochar derived from these straws occurs when the pH
increases from 3 to 7 (Yuan, Xu, & Zhang, 2011). Modification to the bio-
char was also carried out by functionalization with amine-enriched chitosan
over the surface of biochar (Chatterjee et al., 2018). The chelation of lead,
copper and cadmium ions with amine functional group could easily remove
toxic lead through precipitation process. The FT-IR shift in the N-H band
in biochar-chitosan at 3300 cm1 after Pb adsorption is evident in the
interaction between the amine group and Pb ions. The image showing the
germination affected by chitosan modified biochar in the presence and
absence of Pb ions has been presented in Figure 6. It could be clearly iden-
tified that the addition of chitosan-biochar greatly reduces the uptake of
heavy metal by immobilizing the Pb ions in the soil and promotes seed
germination. Such functionalization in biochar will be definitely helpful in
the soil amendment and also the heavy metal risk in the soil would
decrease (Zhou et al., 2013). If we consider the versatility of the soil con-
taining multiple heavy metals, the sorption properties of biochar become
competitive. In this context, the sorption of Pb ions in the presence of Cd
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY 33

Figure 6. Biological effects of dissolved lead (Pb), chitosan-modified biochar (BB-C), and lead-
laden chitosan-modified biochar (BB-C þ Pb) as indicated in the image of seedlings on day 7.
(Zhou et al., 2013).

and Al ions was performed using biochar derived from different feedstocks
from different sources. The adsorption was mainly due to the oxygen-
containing functional groups, phosphates, carbonates and also Si-containing
groups as well. The adsorption of Pb reduces in the presence of Cd, but a
significant reduction is observed in the presence of Al as Al tends to acidify
making biochar counteractive. These factors should be crucially taken into
consideration while working with the soil system (Han et al., 2017).

5.2. Reduction
In perspective with the heavy metal reduction, the multiplicity in the heavy
metals valance helps in reducing the toxicity by complexation or precipita-
tion to convert the toxic ions to intermediate non-toxic ions which are
more stable than their former valence states (Zou et al., 2016). Electron
donating ability of iron/iron oxide nanoparticles embedded in biochar is
quite attractive because the reduction reaction is governed by the redox
potential of Fe(0), which is 0.44 V, facilitating reduction of toxic heavy
metal ions such as Pb2þ/Pb0 (0.13 V), Cu2þ/Cu0 (þ0.34 V) and Cr6þ/
Cr3þ (þ1.36 V) (Liu, Wang, Chen, Zhang, & Lin, 2015). Secondly, the
presence of Fe2þ ions will significantly reduce heavy metal toxicity (Li
et al., 2015). In addition, some surface functional groups such as C-H and
C-OH exhibit certain reduction reactivity. In one of the studies by Zhuang
et al. the reduction of hexavalent chromium to trivalent chromium is due
to the possible oxidation of functional groups C ¼ O or -COOH present on
the biochar surface (Zhuang, Li, Chen, Ma, & Chen, 2014). In addition,
carboxyl functional groups in biochar help in the stabilization of the heavy
metals in the soil as discussed by Uchiyama et al. The biochar rich in carb-
oxyl groups were derived from cottonseed hull and flax shive. In a com-
parison of oxidized and non-oxidized biochar, the former was able to
stabilize Pb, Cu, and Zn, but contradictorily desorbed Sb due to the impact
of the same carboxyl groups affecting the solubility of ions (Uchimiya,
34 S. MANDAL ET AL.

Bannon, & Wartelle, 2012). The carboxyl rich recalcitrant biochar can be
produced from biochar oxidants for high stabilization of heavy metals in
the long-term. Additionally, nano-iron loaded biochar (nZVI@BC) was pre-
pared in a mean to develop low-cost and efficient sorbent (Su, Fang,
Tsang, Zheng, et al., 2016). The results revealed that immobilization of
hexavalent chromium and the total chromium was about 100% and 91%,
respectively for the soil remediation and with optimum dose of 8 g/kg, the
zero-valent iron in the composite converts the existing chromium in
the soil to Cr(OH)3, or oxides/hydroxides of Cr(III)/Fe(III). In addition,
the increase in soil pH due to biochar introduction results in more strong
binding of Fe-Mn oxides, which are not readily available for further proc-
esses in soil. In summary, there is still ambiguity in the occurrence of con-
ditioning of soil via immobilization or by mobilizing the heavy metals.

5.3. Ion-exchange/complexation
Ion-exchange is another mechanism for heavy metals immobilization in the
biochar and its composites through ionization. In general, ion exchange is
via replacement of the ions on the biochar/functionalized biochar surface
with the toxic metal ions. Ion-exchange is expected when functional groups
in the surface ionize at low pH values and surface ions (e.g., Naþ, Kþ,
Mg2þ, and Ca2þ) are exchanged with heavy metal ions. Another study devel-
oped magnetic biochar-MnFe2O4 nanocomposite (BC/FM) using corn straw
and egg white as a precursor for biochar and ferrite pyrolysis, respectively
due to –NH2 and –COOH functional groups. The removal of Pb(II) and
Cd(II) was efficient due to these functional groups and MnFe2O4 dispersion
in biochar. The adsorption capacity with BC/FM for Pb and Cd is about
154.94 and 127.83 mg/g, respectively in the individual-solute system. Pb
adsorption is more favored over Cd in a bi-solute system and the mechanism
that governs apart from Pb/Cd-O complexation is the ion-exchange process.
The biochar contributes to ion-exchange and functional group complexation
by carboxyl and hydroxyl groups, whereas the only complexation is observed
with MnFe2O4 as given in equations below. Also, lower pH is not favorable
for heavy metal binding because of competition from Hþ ions in the system
and magnetism eases the separation (Zhang, Guo, et al., 2019).
High biochar pH value increase soil pH and decrease the mobility of Pb
in contaminated soils. The significant mechanism for Pb immobilization is
the interaction of Pb2þ ions with other cations (e.g., Ca2þ, Kþ, Naþ, Mg2þ)
in biochar. Ion exchange improves the binding of Pb to biochar surfaces
and reduces its mobility significantly. In addition, Pb precipitation on bio-
char surface is known as an essential mechanism of Pb immobilization
with a specific anion (i.e., PO43, OH, Cl) and stable Pb forming
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY 35

compounds, such as hydroxy pyromorphite (Ho, Chen, et al., 2017). In


addition, Wang et al. also noted surface complexation as one of the key
mechanisms of Pb immobilization in the soil through the use of functional
groups (e.g., -OH, -C-H, -C ¼ O,- C ¼ C and –COOH) present in biochar.
Furthermore, the electrostatic attraction of biochar and Pb also has a sig-
nificant impact on the immobilization of soil (Igalavithana et al., 2019).
2R1 COO 
ðSÞ þ MðaqÞ $ ½ðE1 COO Þ2 M ðSÞ
2þ 2þ

2R1 O 
ðSÞ þ MðaqÞ $ ½ðR1 O Þ2 M S
2þ 2þ

2R2 OHðSÞ þ MðaqÞ



$ ½ðR2 O Þ2 M2þ ðSÞ þ 2H þ

Whereas, R1 and R2 are BC and MnFe2O4.


A strong electrostatic attraction can also develop between the charged sur-
face of biochar and the toxic metal ions. When pH is higher than the zero
point charge (pHpzc) of a nano-composite, electrostatic interactions are typ-
ically observed and the negative surface charge is generated by functional
oxygen-containing groups producing positive metal ions. For example, the
biochar modified with manganese oxide nanoparticles could remove about
eight times of Pb2þ ions than the pure biochar because of strong electrostatic
interaction from the surface hydroxyl groups which makes the surface of bio-
char negatively charged (Wang, Sheng, et al., 2015). Strategies like chemical
activation and microwave activation were also conducted for biochar prop-
erty improvement (Liu, Liu, et al., 2018). The relationship between oxygen-
containing groups and positive metal ions is usually linked to complexation.
Another example showed the considerable elimination of arsenic from
the soil, biochar, and Fe-Mn modified biochar composite (FMBCs). It was
concluded that the above changes were due to high levels (e.g., iron oxide,
free manganese oxide, amorphous iron oxide, and amorphous manganese
oxide) in the soil. Additionally, the complexation of soil organic matter
(SOM) could promote heavy metal fixation. The studies showed that differ-
ent forms of arsenic could be transformed under suitable conditions, and
this could happen because of Fe in the FMBC which has provided adsorp-
tion sites and exhibited certain oxidizing ability. In-spite, presence of Mn
benefited particularly to oxidize As(III) to As(V) (Igalavithana et al., 2019).
Similarly, the moieties like graphene and other carbons (CNTs, MWCNTs)
enhances the electron transfer process in the biochar (Wang, Zhao, et al.,
2019). There are also studies on the use of graphene, polymer materials,
and other biochar composites to further explore other possibilities (Abdul,
Zhu, & Chen, 2017).
Concisely, in the practical application of biochar composites, the factors
considered are much more complex than those considered by the labora-
tory. Since, in practice, there will be a wide variety of biological and
36 S. MANDAL ET AL.

microbial communities in the environment, as well as their functionality


varies with climatic conditions. For better understanding of the real-time
effect and changes in the mechanism in soil-based systems, studies on bio-
char based composites need to be developed.

6. Future outlooks and summary


This review provides positive evidence for the implementation of biochar
and biochar based composites in the soil applications. The material prospect
and research of water treatment applications are reported in large quantities,
but the study on soil application of biochar composites is missing to a great
extent. Larger properties and production implementations, as well as the
inclusion of different types of soils, especially in tropical agricultural soils,
are innovative ways to improve crop growth, soil health, and microbial abun-
dance. Consideration has been given to the long-term effects of biochar com-
posites and it is observed that the application of biochar and its composites
is beneficial to enhance soil health and immobilize heavy metals in soil.
However, in order to meet demand and application-based results, there are
still a number of limitations and challenges, as follows:

a. The bare biochar has been a key functional material for a long span.
Although their benefits are well approached and discussed, for enhance-
ment with superior property and longevity with novel materials, such as
nanometals, metal oxides, chemical additives, functional groups, and
graphitization agents need to be studied.
b. To enhance the economic availability, the easier production approach,
biomass feedstock database, and properties influence studies need to be
designed, which could provide a platform to increase the performance
efficiency and decrease the economic burdens.
c. To study and consider the outcome’s as potential and secondary pollu-
tion (back pollution) while applying biochar based composites to the
soil during remediation, the development of more “green synthesis and
fabrication methods” is required.
d. Short term and long-term influence of biochar and biochar based com-
posites on soil physico-chemical properties such as aggregation, pH,
surface potential, the magnitude of density and its influence in soil with
the application need to be ascertained.
e. The short time and long term influence on microbial abundance and
community structure with biochar and biochar based composites within
its porosity, surface potential, pH, SOM and DOC need to be inter-
linked with aging properties of biochar.
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY 37

f. The regeneration of used and field implemented biochar and biochar


based composites and its reusability, change in surface and efficiency
properties, further application in soil, re-implementation perspectives
are also needed to be observed.
g. The considerate of biochar and biochar based composites induced
changes in soil properties should be further developed by compared
laboratory test results (For example, pot experiments, greenhouse
experiments, plant, and soil gene evaluation through real-time quantita-
tive polymerase chain reaction studies (qPCR), and metagenomics) with
the real-time based case studies. Such a comparative evaluation will
help to underline the benefits of biochar composites and bare biochar.
h. The study of the interaction between biochar-based composites with
soil microbes and a better understanding of the toxicity underlying
implementations are necessary to ensure that they are safe for environ-
mental applications and to make the practicality in the field
and commercial.

The production of biochar-based composites can be considered as new


material and provides the advantages of biochar and nanomaterials/metal
oxides. The resulting materials usually have significant improvements in a
functional group, pore properties, surface-active site, catalytic decompos-
ition capacity, ease of separation. Biochar and its composite materials could
also help to improve the soil properties, increase longevity, water retention
and continuously reduce erosion and evaporation in the aerial parts of the
soil and treat soil contaminants as compared to bare biochar’s. Hence,
these materials could provide a distinctive opportunity to be studied and
utilized for various environmental applications.

Acknowledgments
We wish to express our gratitude to the editor and the anonymous reviewers for in-depth
comments, suggestions, and corrections, which will help to greatly improve the manuscript.
The authors thank the peers for their contributions to the works cited here.

Competing interests
The authors declare that they have no competing interests.

Funding
This work was supported by the National Natural Science Foundation of China
(41772264), Applied Basic Research Programs of Science and Technology, Sichuan
Province (18YYJC1745) and the Research Fund of State Key Laboratory of Geohazard
Prevention and Geoenvironment Protection (SKLGP2018Z001).
38 S. MANDAL ET AL.

ORCID
Sandip Mandal http://orcid.org/0000-0002-4535-2048

References
Abbas, Z., Ali, S., Rizwan, M., Zaheer, I. E., Malik, A., Riaz, M. A., … Al-Wabel, M. I.
(2018). A critical review of mechanisms involved in the adsorption of organic and inor-
ganic contaminants through biochar. Arabian Journal of Geosciences, 11, 448. doi:10.
1007/s12517-018-3790-1
Abdul, G., Zhu, X., & Chen, B. (2017). Structural characteristics of biochar-graphene nano-
sheet composites and their adsorption performance for phthalic acid esters. Chemical
Engineering Journal, 319, 9–20. doi:10.1016/j.cej.2017.02.074
Accardi-Dey, A. M., & Gschwend, P. M. (2002). Assessing the combined roles of natural
organic matter and black carbon as sorbents in sediments. Environmental Science &
Technology, 36, 21–29. doi:10.1021/es010953c
Agegnehu, G., Srivastava, A. K., & Bird, M. I. (2017). The role of biochar and biochar-com-
post in improving soil quality and crop performance: A review. Applied Soil Ecology,
119, 156–170. doi:10.1016/j.apsoil.2017.06.008
Agrafioti, E., Kalderis, D., & Diamadopoulos, E. (2014). Arsenic and chromium removal
from water using biochars derived from rice husk, organic solid wastes and sewage
sludge. Journal of Environmental Management, 133, 309–314. doi:10.1016/j.jenvman.2013.
12.007
Ahmad, M., Lee, S. S., Lee, S. E., Al-Wabel, M. I., Tsang, D. C. W., & Ok, Y. S. (2017).
Biochar-induced changes in soil properties affected immobilization/mobilization of met-
als/metalloids in contaminated soils. Journal of Soils and Sediments, 17(3), 717–730. doi:
10.1007/s11368-015-1339-4
Ahmed, M. B., Zhou, J. L., Ngo, H. H., Guo, W., & Chen, M. (2016). Progress in the prep-
aration and application of modified biochar for improved contaminant removal from
water and wastewater. Bioresource Technology, 214, 836–851. doi:10.1016/j.biortech.2016.
05.057
Alaboudi, K. A., Ahmed, B., & Brodie, G. (2019). Effect of biochar on Pb, Cd and Cr avail-
ability and maize growth in artificial contaminated soil. Annals of Agricultural Sciences,
64(1), 95–102. doi:10.1016/j.aoas.2019.04.002
Allen, D. E., Singh, B. P., & Dalal, R. C. (2011). Soil health indicators under climate
change: A review of current knowledge. In B. Singh, A. Cowie, & K. Chan (Eds.), Soil
health and climate change. Soil biology (Vol. 29, pp. 25–45). Berlin, Heidelberg: Springer.
doi:10.1007/978-3-642-20256-8_2
Arabyarmohammadi, H., Darban, A. K., van der Zee, S. E. A. T. M., Abdollahy, M., &
Ayati, B. (2018). Fractionation and leaching of heavy metals in soils amended with a
new biochar nanocomposite. Environmental Science and Pollution Research, 25(7),
6826–6837. doi:10.1007/s11356-017-0976-0
Arabyarmohammadi, H., Khodadadi-Darban, A., Abdollahy, M., Yong, R., Ayati, B.,
Zirakjou, A., & van der Zee, S. (2018). Utilization of a novel chitosan/clay/biochar nano-
biocomposite for immobilization of heavy metals in acid soil environment. Journal of
Polymers and the Environment, 26(5), 2107–2119. doi:10.1007/s10924-017-1102-6
Aung, A., Han, S. H., Youn, W. B., Meng, L., Cho, M. S., & Park, B. B. (2018). Biochar
effects on the seedling quality of Quercus serrata and Prunus sargentii in a containerized
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY 39

production system. Forest Science and Technology, 14(3), 112–118. doi:10.1080/21580103.


2018.1471011
Beckers, F., Mothes, S., Abrigata, J., Zhao, J., Gao, Y., & Rinklebe, J. (2019). Mobilization of
mercury species under dynamic laboratory redox conditions in a contaminated flood-
plain soil as affected by biochar and sugar beet factory lime. Science of the Total
Environment, 672, 604–617. doi:10.1016/j.scitotenv.2019.03.401
Benedetti, V., Patuzzi, F., & Baratieri, M. (2017). Gasification char as a potential substitute
of activated carbon in adsorption applications. Energy Procedia, 105, 712–717. doi:10.
1016/j.egypro.2017.03.380
Berihun, T., Tadele, M., & Kebede, F. (2017). The application of biochar on soil acidity and
other physico-chemical properties of soils in southern Ethiopia. Zeitschrift Fur
Pflanzenernahrung Und Bodenkunde, 180(3), 381–388. doi:10.1002/jpln.201600343
Bogusz, A., & Oleszczuk, P. (2018). Sequential extraction of nickel and zinc in sewage
sludge- or biochar/sewage sludge-amended soil. Science of the Total Environment, 636,
927–935. doi:10.1016/j.scitotenv.2018.04.072
Bombuwala Dewage, N., Fowler, R. E., Pittman, C. U., Mohan, D., & Mlsna, T. (2018).
Lead (Pb2þ) sorptive removal using chitosan-modified biochar: Batch and fixed-bed
studies. RSC Advances, 8(45), 25368–25377. doi:10.1039/C8RA04600J
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. doi:10.1016/j.geo-
derma.2016.07.019
Cao, X., Ma, L., Liang, Y., Gao, B., & Harris, W. (2011). Simultaneous immobilization of
lead and atrazine in contaminated soils using dairy-manure biochar. Environmental
Science & Technology, 45(11), 4884–4889. doi:10.1021/es103752u
Cha, J. S., Park, S. H., Jung, S. C., Ryu, C., Jeon, J. K., Shin, M. C., & Park, Y. K. (2016).
Production and utilization of biochar: A review. Journal of Industrial and Engineering
Chemistry, 40, 1–15. doi:10.1016/j.jiec.2016.06.002
Chatterjee, R., Sajjadi, B., Mattern, D. L., Chen, W.-Y., Zubatiuk, T., Leszczynska, D., …
Hammer, N. (2018). Ultrasound cavitation intensified amine functionalization: A feasible
strategy for enhancing CO2 capture capacity of biochar. Fuel, 225, 287–298. doi:10.1016/
j.fuel.2018.03.145
Chen, C.-W., Kao, C.-M., Hung, C.-M., Chien, C.-C., & Dong, C.-D. (2018). Wood-bio-
char-supported magnetite nanoparticles for remediation of PAH-contaminated estuary
sediment. Catalysts, 8(2), 73. doi:10.3390/catal8020073
Chen, S. S., Yu, I. K. M., Cho, D. W., Song, H., Tsang, D. C. W., Tessonnier, J. P., …
Poon, C. S. (2018). Selective glucose isomerization to fructose via a nitrogen-doped solid
base catalyst derived from spent coffee grounds. ACS Sustainable Chemistry and
Engineering, 6(12), 16113–16120. doi:10.1021/acssuschemeng.8b02752
Chintala, R., Mollinedo, J., Schumacher, T. E., Malo, D. D., & Julson, J. L. (2014). Effect of
biochar on chemical properties of acidic soil. Archives of Agronomy and Soil Science,
60(3), 393–404. doi:10.1080/03650340.2013.789870
Cho, D.-W., Yoon, K., Ahn, Y., Sun, Y., Tsang, D. C. W., Hou, D., … Song, H. (2019).
Fabrication and environmental applications of multifunctional mixed metal-biochar com-
posites (MMBC) from red mud and lignin wastes. Journal of Hazardous Materials, 374,
412–419. doi:10.1016/j.jhazmat.2019.04.071
Christianson, L., Hedley, M., Camps, M., Free, H., & Saggar, S. (2010). Influence of biochar
amendments on denitrification bioreactor performance. Methods, 1–8.
40 S. MANDAL ET AL.

Cui, J., Jin, Q., Li, Y., & Li, F. (2019). Oxidation and removal of As(iii) from soil using
novel magnetic nanocomposite derived from biomass waste. Environmental Science:
Nano, 6, 478–488. doi:10.1039/C8EN01257A
Cybulak, M., Sokołowska, Z., Boguta, P., & Tomczyk, A. (2019). Influence of pH and grain
size on physicochemical properties of biochar and released humic substances. Fuel, 240,
334–338. doi:10.1016/j.fuel.2018.12.003
Devi, P., & Saroha, A. K. (2014). Synthesis of the magnetic biochar composites for use as
an adsorbent for the removal of pentachlorophenol from the effluent. Bioresource
Technology, 169, 525–531. doi:10.1016/j.biortech.2014.07.062
Devi, P., & Saroha, A. K. (2015a). Effect of pyrolysis temperature on polycyclic aromatic
hydrocarbons toxicity and sorption behaviour of biochars prepared by pyrolysis of paper
mill effluent treatment plant sludge. Bioresource Technology, 192, 312–320. doi:10.1016/j.
biortech.2015.05.084
Devi, P., & Saroha, A. K. (2015b). Simultaneous adsorption and dechlorination of penta-
chlorophenol from effluent by Ni-ZVI magnetic biochar composites synthesized from
paper mill sludge. Chemical Engineering Journal, 271, 195–203. doi:10.1016/j.cej.2015.02.
087
Doran, J. W. (2002). Soil health and global sustainability: Translating science into practice.
Agriculture, Ecosystems and Environment, 88(2), 119–127. doi:10.1016/S0167-8809(01)00246-
8
Doran, J. W., Sarrantonio, M., & Liebig, M. A. (1996). Soil health and sustainability.
Advances in Agronomy, 56, 1–229. doi:10.1016/S0065-2113(08)60178-9
Doran, J. W., & Zeiss, M. R. (2000). Soil health and sustainability: Managing the biotic
component of soil quality. Applied Soil Ecology, 15(1), 3–11. doi:10.1016/S0929-
1393(00)00067-6
Elzobair, K. A., Stromberger, M. E., Ippolito, J. A., & Lentz, R. D. (2016). Contrasting
effects of biochar versus manure on soil microbial communities and enzyme activities in
an Aridisol. Chemosphere, 142, 145–152. doi:10.1016/j.chemosphere.2015.06.044
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. doi:10.1016/j.biortech.2012.03.022
Fahmi, A. H., Samsuri, A. W., Jol, H., & Singh, D. (2018). Physical modification of biochar
to expose the inner pores and their functional groups to enhance lead adsorption. RSC
Advances, 8(67), 38270–38280. doi:10.1039/C8RA06867D
Fang, B., Li, X., & Bin, Z. (2014). Influence of biochar on soil physical and chemical prop-
erties and crop yields in rainfed field. Journal of Ecology and Environmental Sciences,
23(8), 1292–1297.
Foster, E. J., Hansen, N., Wallenstein, M., & Cotrufo, M. F. (2016). Biochar and manure
amendments impact soil nutrients and microbial enzymatic activities in a semi-arid irri-
gated maize cropping system. Agriculture, Ecosystems and Environment, 233, 404–414.
doi:10.1016/j.agee.2016.09.029
Gai, X-P., & Zhai, L-M. W. H. (2017). Impacts of biochar on soil microbial biomass and
community structure. Journal of Shenyang Agricultural University, 48(4), 399–410. doi:10.
1016/j.ejsobi.2016.02.004
Gale, W. J., Cambardella, C. A., & Bailey, T. B. (2000). Root-derived carbon and the forma-
tion and stabilization of aggregates. Soil Science Society of America Journal, 64(1),
201–207. doi:10.2136/sssaj2000.641201x
Gao, X., Peng, Y., Zhou, Y., Adeel, M., & Chen, Q. (2019). Effects of magnesium ferrite
biochar on the cadmium passivation in acidic soil and bioavailability for packoi (Brassica
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY 41

chinensis L.). Journal of Environmental Management, 251, 109610. doi:10.1016/j.jenvman.


2019.109610
Gaskin, J. W., Steiner, C., Harris, K., Das, K. C., & Bibens, B. (2008). Effect of low-tem-
perature pyrolysis conditions on biochar for agricultural use. Transactions of the ASABE,
51(6), 2061–2069.
Gaunt, J., & Cowie, A. (2012). Biochar, greenhouse gas accounting and emissions trading.
In J. Lehmann & S. Joseph (Eds.), Biochar for environmental management: Science and
technology (p. 24). New York: Routledge. doi:10.4324/9781849770552
Geroy, I. J., Gribb, M. M., Marshall, H. P., Chandler, D. G., Benner, S. G., & Mcnamara,
J. P. (2011). Aspect influences on soil water retention and storage. Hydrological Processes,
25(25), 3836–3842. doi:10.1002/hyp.8281
Githinji, L. (2014). Effect of biochar application rate on soil physical and hydraulic proper-
ties of a sandy loam. Archives of Agronomy and Soil Science, 60(4), 457–470. doi:10.1080/
03650340.2013.821698
Gła˛b, T., Palmowska, J., Zaleski, T., & Gondek, K. (2016). Effect of biochar application on
soil hydrological properties and physical quality of sandy soil. Geoderma, 281, 11–20.
doi:10.1016/j.geoderma.2016.06.028
Gomes, H. I., Dias-Ferreira, C., & Ribeiro, A. B. (2012). Electrokinetic enhanced transport
of zero valent iron nanoparticles for chromium (VI) reduction in soils. Chemical
Engineering Transactions, 28, 139–144. doi:10.3303/CET1228024
Gonzaga, M. I. S., Mackowiak, C., Quint~ao de Almeida, A., Wisniewski, A., Figueiredo de
Souza, D., da Silva Lima, I., & Nascimento de Jesus, A. (2018). Assessing biochar appli-
cations and repeated Brassica juncea L. production cycles to remediate Cu contaminated
soil. Chemosphere, 201, 278–285. doi:10.1016/j.chemosphere.2018.03.038
Goodman, A. M., & Ennos, A. R. (2007). A comparative study of the response of the roots
and shoots of sunflower and maize to mechanical stimulation. Journal of Experimental
Botany, 47(10), 1499–1507. doi:10.1093/jxb/47.10.1499
Gul, S., Whalen, J. K., Thomas, B. W., Sachdeva, V., & Deng, H. (2015). Physico-chemical
properties and microbial responses in biochar-amended soils: Mechanisms and future
directions. Agriculture, Ecosystems and Environment, 206, 46–59. doi:10.1016/j.agee.2015.
03.015
Han, L., Qian, L., Liu, R., Chen, M., Yan, J., & Hu, Q. (2017). Lead adsorption by biochar
under the elevated competition of cadmium and aluminum. Scientific Reports, 7(1), 2264.
doi:10.1038/s41598-017-02353-4
Han, Z., Sani, B., Mrozik, W., Obst, M., Beckingham, B., Karapanagioti, H. K., & Werner,
D. (2015). Magnetite impregnation effects on the sorbent properties of activated carbons
and biochars. Water Research, 70, 394–403. doi:10.1016/j.watres.2014.12.016
Hansen, V., M€ uller-St€
over, D., Munkholm, L. J., Peltre, C., Hauggaard-Nielsen, H., &
Jensen, L. S. (2016). The effect of straw and wood gasification biochar on carbon seques-
tration, selected soil fertility indicators and functional groups in soil: An incubation
study. Geoderma, 269, 99–107. doi:10.1016/j.geoderma.2016.01.033
He, L. L., Zhong, Z. K., & Yang, H. M. (2017). Effects on soil quality of biochar and straw
amendment in conjunction with chemical fertilizers. Journal of Integrative Agriculture,
16(3), 704–712. doi:10.1016/S2095-3119(16)61420-X
Heitk€otter, J., & Marschner, B. (2015). Interactive effects of biochar ageing in soils related
to feedstock, pyrolysis temperature, and historic charcoal production. Geoderma,
245–246, 56–64. doi:10.1016/j.geoderma.2015.01.012
42 S. MANDAL ET AL.

Herath, H. M. S. K., Camps-Arbestain, M., & Hedley, M. (2013). Effect of biochar on soil
physical properties in two contrasting soils: An Alfisol and an Andisol. Geoderma,
209–210, 188–197. doi:10.1016/j.geoderma.2013.06.016
Ho, S. H., Chen, Y. D., Yang, Z. K., Nagarajan, D., Chang, J. S., & Ren, N. Q. (2017).
High-efficiency removal of lead from wastewater by biochar derived from anaerobic
digestion sludge. Bioresource Technology, 246, 142–149. doi:10.1016/j.biortech.2017.08.025
Ho, S. H., Zhu, S., & Chang, J. S. (2017). Recent advances in nanoscale-metal assisted bio-
char derived from waste biomass used for heavy metals removal. Bioresource Technology,
246, 123–134. doi:10.1016/j.biortech.2017.08.061
Houben, D., Evrard, L., & Sonnet, P. (2013). Mobility, bioavailability and pH-dependent
leaching of cadmium, zinc and lead in a contaminated soil amended with biochar.
Chemosphere, 92(11), 1450–1457. doi:10.1016/j.chemosphere.2013.03.055
Huang, B. C., Jiang, J., Huang, G. X., & Yu, H. Q. (2018). Sludge biochar-based catalysts
for improved pollutant degradation by activating peroxymonosulfate. Journal of
Materials Chemistry A, 6, 8978–8985. doi:10.1039/C8TA02282H
Huang, C., Liu, S., Li, R., Sun, F., Zhou, Y., & Yu, G. (2016). Spectroscopic evidence of the
improvement of reactive iron mineral content in red soil by long-term application of
swine manure. PLoS One, 11(1), e0146364. doi:10.1371/journal.pone.0146364
Huang, M., Fan, L., Jiang, L., Yang, S., Zou, Y., & Uphoff, N. (2019). Continuous applica-
tions of biochar to rice: Effects on grain yield and yield attributes. Journal of Integrative
Agriculture, 18(3), 563–570. doi:10.1016/S2095-3119(18)61993-8
Igalavithana, A. D., Kwon, E. E., Vithanage, M., Rinklebe, J., Moon, D. H., Meers, E., …
Ok, Y. S. (2019). Soil lead immobilization by biochars in short-term laboratory incuba-
tion studies. Environment International, 127, 190–198. doi:10.1016/j.envint.2019.03.031
Igalavithana, A. D., Lee, S. E., Lee, Y. H., Tsang, D. C. W., Rinklebe, J., Kwon, E. E., & Ok,
Y. S. (2017). Heavy metal immobilization and microbial community abundance by vege-
table waste and pine cone biochar of agricultural soils. Chemosphere, 174, 593–603. doi:
10.1016/j.chemosphere.2017.01.148
Inyang, M., Gao, B., Zimmerman, A., Zhou, Y., & Cao, X. (2015). Sorption and cosorption
of lead and sulfapyridine on carbon nanotube-modified biochars. Environmental Science
and Pollution Research, 22(3), 1868–1876. doi:10.1007/s11356-014-2740-z
Jeffery, S., Verheijen, F. G. A., van der Velde, M., & Bastos, A. C. (2011). A quantitative
review of the effects of biochar application to soils on crop productivity using meta-ana-
lysis. Agriculture, Ecosystems and Environment, 144(1), 175–187. doi:10.1016/j.agee.2011.
08.015
Jien, S. H., & Wang, C. S. (2013). Effects of biochar on soil properties and erosion potential
in a highly weathered soil. CATENA, 110, 225–233. doi:10.1016/j.catena.2013.06.021
Keiluweit, M., Nico, P. S., Johnson, M. G., & Kleber, M. (2010). Dynamic molecular struc-
ture of plant biomass-derived black carbon (biochar). Environmental Science &
Technology, 44(4), 1247–1253. doi:10.1021/es9031419
Kim, H.-B., Kim, S.-H., Jeon, E.-K., Kim, D.-H., Tsang, D. C. W., Alessi, D. S., … Baek,
K. (2018). Effect of dissolved organic carbon from sludge, rice straw and spent coffee
ground biochar on the mobility of arsenic in soil. Science of the Total Environment, 636,
1241–1248. doi:10.1016/j.scitotenv.2018.04.406
Krishnakumar, S., Rajalakshmi, A. G., Balaganesh, B., Manikandan, P., Vinoth, C., &
Rajendran, V. (2014). Impact of biochar on soil physical properties. International Journal
of Advanced Research, 4(5), 280–284.
Ladygina, N., & Rineau, F. (2013). Biochar and soil biota. Boca Raton, FL: CRC Press.
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY 43

Laishram, J., Saxena, K. G., Maikhuri, R. K., & Rao, K. S. (2012). Soil quality and soil
health: A review. International Journal of Ecology and Environmental Sciences, 38(1),
19–37.
Lal, R. (2009). Soils and sustainable agriculture: A review. Sustainable Agriculture, 28(1),
57–64. doi:10.1007/978-90-481-2666-8_3
Lee, S.-H., Kim, S.-B., Choi, J.-W., Chung, S.-G., & Mahendran, B. (2014). Surface modified
mesostructured iron oxyhydroxide: Synthesis, ecotoxicity, and application. Water
Environment Research, 86(12), 2338–2346. doi:10.2175/106143014X14062131178790
Lehmann, J., & Joseph, S. (2012a). Biochar for environmental management: An introduc-
tion. In J. Lehmann & S. Joseph (Eds.), Biochar for environmental management: Science
and technology (pp. 1–12). London: Routledge.
Lehmann, J., & Joseph, S. (2012b). Biochar systems. In J. Lehmann & S. Joseph (Eds.),
Biochar for environmental management: Science and technology (pp. 1–22). London:
Routledge.
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. doi:10.1016/j.soilbio.2011.04.022
Li, G., Khan, S., Ibrahim, M., Sun, T. R., Tang, J. F., Cotner, J. B., & Xu, Y. Y. (2018).
Biochars induced modification of dissolved organic matter (DOM) in soil and its impact
on mobility and bioaccumulation of arsenic and cadmium. Journal of Hazardous
Materials, 348, 100–108. doi:10.1016/j.jhazmat.2018.01.031
Li, H., Xu, H., Zhou, S., Yu, Y., Li, H., Zhou, C., … Wang, G. (2018). Distribution and
transformation of lead in rice plants grown in contaminated soil amended with biochar
and lime. Ecotoxicology and Environmental Safety, 165, 589–596. doi:10.1016/j.ecoenv.
2018.09.039
Li, L., Zhang, K., Chen, L., Huang, Z., Liu, G., Li, M., & Wen, Y. (2017). Mass preparation
of micro/nano-powders of biochar with water-dispersibility and their potential applica-
tion. New Journal of Chemistry, 41, 9649–9657. doi:10.1039/C7NJ00742F
Li, M., Liu, M., Li, Z. P., Jiang, C. Y., & Wu, M. (2016). Soil N transformation and micro-
bial community structure as affected by adding biochar to a paddy soil of subtropical
China. Journal of Integrative Agriculture, 15(1), 209–219. doi:10.1016/S2095-
3119(15)61136-4
Li, X., Zhang, W., Liu, T., Chen, L., Chen, P., & Li, F. (2016). Changes in the composition
and diversity of microbial communities during anaerobic nitrate reduction and Fe(II)
oxidation at circumneutral pH in paddy soil. Soil Biology and Biochemistry, 94, 70–79.
doi:10.1016/j.soilbio.2015.11.013
Li, Y., Cheng, W., Sheng, G., Li, J., Dong, H., Chen, Y., & Zhu, L. (2015). Synergetic effect
of a pillared bentonite support on SE(VI) removal by nanoscale zero valent iron. Applied
Catalysis B: Environmental, 174–175, 329–335. doi:10.1016/j.apcatb.2015.03.025
Li, Y., Hu, S., Chen, J., M€ uller, K., Li, Y., Fu, W., … Wang, H. (2018). Effects of biochar
application in forest ecosystems on soil properties and greenhouse gas emissions: A
review. Journal of Soils and Sediments, 18(2), 546–563. doi:10.1007/s11368-017-1906-y
Lian, F., & Xing, B. (2017). Black carbon (biochar) in water/soil environments: Molecular
structure, sorption, stability, and potential risk. Environmental Science and Technology,
51, 13517–13532. doi:10.1021/acs.est.7b02528
Lin, L., Gao, M., Qiu, W., Wang, D., Huang, Q., & Song, Z. (2017). Reduced arsenic accu-
mulation in indica rice (Oryza sativa L.) cultivar with ferromanganese oxide impregnated
biochar composites amendments. Environmental Pollution, 231, 479–486. doi:10.1016/j.
envpol.2017.08.001
44 S. MANDAL ET AL.

Lin, L., Li, Z., Liu, X., Qiu, W., & Song, Z. (2019). Effects of Fe-Mn modified biochar com-
posite treatment on the properties of As-polluted paddy soil. Environmental Pollution,
244, 600–607. doi:10.1016/j.envpol.2018.10.011
Ling, N., Zhu, C., Xue, C., Chen, H., Duan, Y., Peng, C., … Shen, Q. (2016). Insight into
how organic amendments can shape the soil microbiome in long-term field experiments
as revealed by network analysis. Soil Biology and Biochemistry, 99, 137–149. doi:10.1016/
j.soilbio.2016.05.005
Liu, M., Wang, Y., Chen, L., Zhang, Y., & Lin, Z. (2015). Mg(OH)2 supported nanoscale
zero valent iron enhancing the removal of Pb(II) from aqueous solution. ACS Applied
Materials and Interfaces, 7, 7961–7969. doi:10.1021/am509184e
Liu, S. J., Liu, Y. G., Tan, X. F., Zeng, G. M., Zhou, Y. h., Liu, S. B., … Wen, J. (2018).
The effect of several activated biochars on Cd immobilization and microbial community
composition during in-situ remediation of heavy metal contaminated sediment.
Chemosphere, 208, 655–664. doi:10.1016/j.chemosphere.2018.06.023
Liu, T., Gao, B., Fang, J., Wang, B., & Cao, X. (2016). Biochar-supported carbon nanotube
and graphene oxide nanocomposites for Pb(ii) and Cd(ii) removal. RSC Advances, 6(29),
24314–24319. doi:10.1039/C6RA01895E
Liu, X., Zhang, A., Ji, C., Joseph, S., Bian, R., Li, L., … Paz-Ferreiro, J. (2013). Biochar’s
effect on crop productivity and the dependence on experimental conditions - A meta-
analysis of literature data. Plant and Soil, 373(1-2), 583–594. doi:10.1007/s11104-013-
1806-x
Liu, Z., Xu, J., Li, X., & Wang, J. (2018). Mechanisms of biochar effects on thermal proper-
ties of red soil in south China. Geoderma, 323, 41–51. doi:10.1016/j.geoderma.2018.02.
045
Lu, H., Gong, Y., Tang, J., Huang, Y., & Gao, K. (2015). Advances in preparation and
applications of biochar and its composites. Journal of Agro-Environment Science, 6(1),
1–12. doi:10.11654/jaes.2015.08.001
Lyu, H., Gong, Y., Tang, J., Huang, Y., & Wang, Q. (2016). Immobilization of heavy metals
in electroplating sludge by biochar and iron sulfide. Environmental Science and Pollution
Research, 23(14), 14472–14488. doi:10.1007/s11356-016-6621-5
Lyu, H., He, Y., Tang, J., Hecker, M., Liu, Q., Jones, P. D., … Giesy, J. P. (2016). Effect of
pyrolysis temperature on potential toxicity of biochar if applied to the environment.
Environmental Pollution, 218, 1–7. doi:10.1016/j.envpol.2016.08.014
Lyu, H., Zhao, H., Tang, J., Gong, Y., Huang, Y., Wu, Q., & Gao, B. (2018).
Immobilization of hexavalent chromium in contaminated soils using biochar supported
nanoscale iron sulfide composite. Chemosphere, 194, 360–369. doi:10.1016/j.chemosphere.
2017.11.182
Ma, H., Li, J. B., Liu, W. W., Miao, M., Cheng, B. J., & Zhu, S. W. (2015). Novel synthesis
of a versatile magnetic adsorbent derived from corncob for dye removal. Bioresource
Technology, 190, 13–20. doi:10.1016/j.biortech.2015.04.048
Mahar, A., Wang, P., Li, R., & Zhang, Z. (2015). Immobilization of lead and cadmium in
contaminated soil using amendments: A review. Pedosphere, 25(4), 555–568. doi:10.1016/
S1002-0160(15)30036-9
Mandal, S., Bhattacharya, T. K., Verma, A. K., & Haydary, J. (2018). Optimization of pro-
cess parameters for bio-oil synthesis from pine needles (Pinus roxburghii) using response
surface methodology. Chemical Papers, 72(3), 603–616. doi:10.1007/s11696-017-0306-5
Marıa, E., Gonzalez, M., Romero-Hermoso Osorio, L., Gonzalez, A., Hidalgo, P., Meier, S.,
… Cea, M. (2017). Effects of pyrolysis conditions on physicochemical properties of oat
hull derived biochar. BioResources, 12(1), 2040–2057. doi:10.15376/biores.12.1.2040-2057
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY 45

Martinez-Salgado, M. M., Gutierrez-Romero, V., Jannsens, M., & Ortega-Blu, R. (2010).


Biological soil quality indicators: A review. Current Research, Technology and Education
Topics in Applied Microbiology and Microbial Biotechnology, 2010, 319–328.
Mary, B., Recous, S., Darwis, D., & Robin, D. (1996). Interactions between decomposition
of plant residues and nitrogen cycling in soil. Plant and Soil, 181(1), 71–82. doi:10.1007/
BF00011294
Masto, R. E., Kumar, S., Rout, T. K., Sarkar, P., George, J., & Ram, L. C. (2013). Biochar
from water hyacinth (Eichornia crassipes) and its impact on soil biological activity.
CATENA, 111, 64–71. doi:10.1016/j.catena.2013.06.025
Meng, F., Zhu, Y., Li, H., Zhang, W., Wu, S., Zhang, G., & Li, M. (2017). Effects of the
remediation of Cr(VI) in soil by nanoscale zero-valent iron (nZVI) with modified bio-
char. Huanjing Kexue Xuebao/Acta Scientiae Circumstantiae. doi:10.13671/j.hjkxxb.2017.
0240
Mitra, A., Chatterjee, S., Moogouei, R., & Gupta, D. K. (2017). Arsenic accumulation in
rice and probable mitigation approaches: A review. Agronomy, 7(4), 67. doi:10.3390/
agronomy7040067
Mohan, D., Kumar, H., Sarswat, A., Alexandre-Franco, M., & Pittman, C. U. (2014).
Cadmium and lead remediation using magnetic oak wood and oak bark fast pyrolysis
bio-chars. Chemical Engineering Journal, 236, 513–528. doi:10.1016/j.cej.2013.09.057
Mohan, D., Sarswat, A., Ok, Y. S., & Pittman, C. U. (2014). Organic and inorganic contam-
inants removal from water with biochar, a renewable, low cost and sustainable adsorbent
- A critical review. Bioresource Technology, 160, 191–202. doi:10.1016/j.biortech.2014.01.
120
Moon, D. H., Park, J.-W., Chang, Y.-Y., Ok, Y. S., Lee, S. S., Ahmad, M., … Baek, K.
(2013). Immobilization of lead in contaminated firing range soil using biochar.
Environmental Science and Pollution Research, 20(12), 8464–8471. doi:10.1007/s11356-
013-1964-7
Mukherjee, A., Lal, R., & Zimmerman, A. R. (2014). Effects of biochar and other amend-
ments on the physical properties and greenhouse gas emissions of an artificially degraded
soil. Science of the Total Environment, 487, 26–36. doi:10.1016/j.scitotenv.2014.03.141
Munda, S., Bhaduri, D., Mohanty, S., Chatterjee, D., Tripathi, R., Shahid, M., … Nayak,
A. K. (2018). Dynamics of soil organic carbon mineralization and C fractions in paddy
soil on application of rice husk biochar. Biomass and Bioenergy, 115, 1–9. doi:10.1016/j.
biombioe.2018.04.002
Nannipieri, P., Ascher, J., Ceccherini, M. T., Landi, L., Pietramellara, G., & Renella, G.
(2017). Microbial diversity and soil functions. European Journal of Soil Science, 68(1),
12–26. doi:10.1111/ejss.4_12398
Nelissen, V., R€utting, T., Huygens, D., Staelens, J., Ruysschaert, G., & Boeckx, P. (2012).
Maize biochars accelerate short-term soil nitrogen dynamics in a loamy sand soil. Soil
Biology and Biochemistry, 55, 20–27. doi:10.1016/j.soilbio.2012.05.019
O’Connor, D., Peng, T., Li, G., Wang, S., Duan, L., Mulder, J., … Hou, D. (2018). Sulfur-
modified rice husk biochar: A green method for the remediation of mercury contami-
nated soil. Science of the Total Environment, 621, 819–826. doi:10.1016/j.scitotenv.2017.
11.213
Ok, Y. S., Uchimiya, S. M., Chang, S. X., & Bolan, N. (2016). Biochar: Production, charac-
terization, and applications. Boca Raton, FL: CRC Press.
Oliveira, F. R., Patel, A. K., Jaisi, D. P., Adhikari, S., Lu, H., & Khanal, S. K. (2017).
Environmental application of biochar: Current status and perspectives. Bioresource
Technology, 246, 110–122. doi:10.1016/j.biortech.2017.08.122
46 S. MANDAL ET AL.

Pan, D., Liu, C., Yu, H., & Li, F. (2019). A paddy field study of arsenic and cadmium pol-
lution control by using iron-modified biochar and silica sol together. Environmental
Science and Pollution Research, 26(24), 24979–24987. doi:10.1007/s11356-019-05381-x
Park, J. H., Choppala, G. K., Bolan, N. S., Chung, J. W., & Chuasavathi, T. (2011). Biochar
reduces the bioavailability and phytotoxicity of heavy metals. Plant and Soil, 348(1–2),
439–451. doi:10.1007/s11104-011-0948-y
Patil, S. S., Shedbalkar, U. U., Truskewycz, A., Chopade, B. A., & Ball, A. S. (2016).
Nanoparticles for environmental clean-up: A review of potential risks and emerging solu-
tions. Environmental Technology & Innovation, 5, 10–21. doi:10.1016/j.eti.2015.11.001
Patra, J. M., Panda, S. S., & Dhal, N. K. (2017). Biochar as a low-cost adsorbent for heavy
metal removal: A review. International Journal of Research in Biosciences, 6(1), 1–7.
Peng, F., & Sun, R. C. (2010). Modification of cereal straws as natural sorbents for remov-
ing metal ions from industrial waste water. In R. C. Sun (Ed.), Cereal straw as a resource
for sustainable biomaterials and biofuels (pp. 219–237). New York: Elsevier.
Penido, E. S., Martins, G. C., Mendes, T. B. M., Melo, L. C. A., do Rosario Guimar~aes, I.,
& Guilherme, L. R. G. (2019). Combining biochar and sewage sludge for immobilization
of heavy metals in mining soils. Ecotoxicology and Environmental Safety, 172, 326–333.
doi:10.1016/j.ecoenv.2019.01.110
Premarathna, K. S. D., Rajapaksha, A. U., Sarkar, B., Kwon, E. E., Bhatnagar, A., Ok, Y. S.,
& Vithanage, M. (2019). Biochar-based engineered composites for sorptive decontamin-
ation of water: A review. Chemical Engineering Journal, 372, 536–550. doi:10.1016/j.cej.
2019.04.097
Qi, F., Yan, Y., Lamb, D., Naidu, R., Bolan, N. S., Liu, Y., … Semple, K. T. (2017).
Thermal stability of biochar and its effects on cadmium sorption capacity. Bioresource
Technology, 246, 48–56. doi:10.1016/j.biortech.2017.07.033
Qian, K., Kumar, A., Zhang, H., Bellmer, D., & Huhnke, R. (2015). Recent advances in util-
ization of biochar. Renewable and Sustainable Energy Reviews, 42, 1055–1064. doi:10.
1016/j.rser.2014.10.074
Qian, T. T., Wu, P., Qin, Q. Y., Huang, Y. N., Wang, Y. J., & Zhou, D. M. (2019).
Screening of wheat straw biochars for the remediation of soils polluted with Zn (II) and
Cd (II). Journal of Hazardous Materials, 362, 311–317. doi:10.1016/j.jhazmat.2018.09.034
Qian, W., Zhao, A. Z., & Xu, R. K. (2013). Sorption of As(V) by aluminum-modified crop
straw-derived biochars. Water, Air, and Soil Pollution, 224, 1610. doi:10.1007/s11270-
013-1610-5
Qiao, J-T., Liu, T-X., Wang, X-Q., Li, F-B., Lv, Y-h., Cui, J-h., … Liu, C-P. (2018).
Simultaneous alleviation of cadmium and arsenic accumulation in rice by applying zero-
valent iron and biochar to contaminated paddy soils. Chemosphere, 195, 260–271. doi:10.
1016/j.chemosphere.2017.12.081
Qiao, Y., Wu, J., Xu, Y., Fang, Z., Zheng, L., Cheng, W., … Zhao, D. (2017). Remediation
of cadmium in soil by biochar-supported iron phosphate nanoparticles. Ecological
Engineering, 106, 515–522. doi:10.1016/j.ecoleng.2017.06.023
Rafiq, M. K., Joseph, S. D., Li, F., Bai, Y., Shang, Z., Rawal, A., … Long, R.-J. (2017).
Pyrolysis of attapulgite clay blended with yak dung enhances pasture growth and soil
health: Characterization and initial field trials. Science of the Total Environment,
607–608, 184–194. doi:10.1016/j.scitotenv.2017.06.186
Rajapaksha, A. U., Chen, S. S., Tsang, D. C. W., Zhang, M., Vithanage, M., Mandal, S., …
Ok, Y. S. (2016). Engineered/designer biochar for contaminant removal/immobilization
from soil and water: Potential and implication of biochar modification. Chemosphere,
148, 276–291. doi:10.1016/j.chemosphere.2016.01.043
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY 47

Rajendran, M., Shi, L., Wu, C., Li, W., An, W., Liu, Z., & Xue, S. (2019). Effect of sulfur
and sulfur-iron modified biochar on cadmium availability and transfer in the soil–rice
system. Chemosphere, 222, 314–322. doi:10.1016/j.chemosphere.2019.01.149
Ralebitso, T. K., Sr., & Orr, C. H. (2016). Biochar application essential soil microbial ecology.
London: Elsevier. doi:10.1016/B978-0-12-803433-0.00001-1
Reeves, D. W. (1997). The role of soil organic matter in maintaining soil quality in con-
tinuous cropping systems. Soil and Tillage Research, 43(1–2), 131–167. doi:10.1016/
S0167-1987(97)00038-X
Riddle, M., Bergstr€ om, L., Schmieder, F., Lundberg, D., Condron, L., & Cederlund, H.
(2018). Impact of biochar coated with magnesium (hydr)oxide on phosphorus leaching
from organic and mineral soils. Journal of Soils and Sediments, 19, 1875–1889. doi:10.
1007/s11368-018-2197-7
Rizwan, M. S., Imtiaz, M., Huang, G., Chhajro, M. A., Liu, Y., Fu, Q., … Hu, H. (2016).
Immobilization of Pb and Cu in polluted soil by superphosphate, multi-walled carbon
nanotube, rice straw and its derived biochar. Environmental Science and Pollution
Research, 23(15), 15532–15543. doi:10.1007/s11356-016-6695-0
Rocco, C., Seshadri, B., Adamo, P., Bolan, N. S., Mbene, K., & Naidu, R. (2018). Impact of
waste-derived organic and inorganic amendments on the mobility and bioavailability of
arsenic and cadmium in alkaline and acid soils. Environmental Science and Pollution
Research, 25(26), 25896–25905. doi:10.1007/s11356-018-2655-1
Rodriguez, A., Lemos, D., Trujillo, Y. T., Amaya, J. G., & Ramos, L. D. (2019).
Effectiveness of biochar obtained from corncob for immobilization of lead in contami-
nated soil. Journal of Health and Pollution, 9(23), 190907. doi:10.5696/2156-9614-9.23.
190907
Saadat, S., Raei, E., & Talebbeydokhti, N. (2018). Enhanced removal of phosphate from
aqueous solutions using a modified sludge derived biochar: Comparative study of various
modifying cations and RSM based optimization of pyrolysis parameters. Journal of
Environmental Management, 225, 75–83. doi:10.1016/j.jenvman.2018.07.037
Sajjadi, B., Chen, W. Y., & Egiebor, N. O. (2019). A comprehensive review on physical acti-
vation of biochar for energy and environmental applications. Reviews in Chemical
Engineering, 35(6), 735–776. doi:10.1515/revce-2017-0113
Salam, A., Shaheen, S. M., Bashir, S., Khan, I., Wang, J., Rinklebe, J., … Hu, H. (2019).
Rice straw- and rapeseed residue-derived biochars affect the geochemical fractions and
phytoavailability of Cu and Pb to maize in a contaminated soil under different moisture
content. Journal of Environmental Management, 237, 5–14. doi:10.1016/j.jenvman.2019.
02.047
Sandhu, S. S., Ussiri, D. A. N., Kumar, S., Chintala, R., Papiernik, S. K., Malo, D. D., &
Schumacher, T. E. (2017). Analyzing the impacts of three types of biochar on soil carbon
fractions and physiochemical properties in a corn-soybean rotation. Chemosphere, 184,
473–481. doi:10.1016/j.chemosphere.2017.05.165
Santos, M. P. F., Brito, M. J. P., Junior, E. C. S., Bonomo, R. C. F., & Veloso, C. M. (2019).
Pepsin immobilization on biochar by adsorption and covalent binding, and its applica-
tion for hydrolysis of bovine casein. Journal of Chemical Technology & Biotechnology,
94(6), 1982–1990. doi:10.1002/jctb.5981
Schomberg, H. H., Gaskin, J. W., Harris, K., Das, K. C., Novak, J. M., Busscher, W. J., …
Xing, B. (2012). Influence of biochar on nitrogen fractions in a coastal plain soil. Journal
of Environment Quality, 41(4), 1087. doi:10.2134/jeq2011.0133
Sharma, M., Singh, J., Baskar, C., & Kumar, A. (2018). A comprehensive review on biochar
formation and its utilization for wastewater treatment. Pollution Research, 37, S1–S18.
48 S. MANDAL ET AL.

Shen, Z., Zhang, J., Hou, D., Tsang, D. C. W., Ok, Y. S., & Alessi, D. S. (2019). Synthesis
of MgO-coated corncob biochar and its application in lead stabilization in a soil washing
residue. Environment International, 122, 357–362. doi:10.1016/j.envint.2018.11.045
Sheng, Y., Zhan, Y., & Zhu, L. (2016). Reduced carbon sequestration potential of biochar
in acidic soil. Science of the Total Environment, 572, 129–137. doi:10.1016/j.scitotenv.
2016.07.140
Sohi, S. P., Krull, E., Lopez-Capel, E., & Bol, R. (2010). A review of biochar and its use and
function in soil. In Advances in agronomy (Vol. 105, pp. 47–82). Cambridge, MA:
Academic Press. doi:10.1016/S0065-2113(10)05002-9
Steiner, C., Das, K. C., Garcia, M., F€ orster, B., & Zech, W. (2008). Charcoal and smoke
extract stimulate the soil microbial community in a highly weathered xanthic Ferralsol.
Pedobiologia, 51(5–6), 359–366. doi:10.1016/j.pedobi.2007.08.002
Stewart, C. E., Zheng, J., Botte, J., & Cotrufo, M. F. (2013). Co-generated fast pyrolysis bio-
char mitigates green-house gas emissions and increases carbon sequestration in temper-
ate soils. GCB Bioenergy, 5(2), 153–164. doi:10.1111/gcbb.12001
Su, H., Fang, Z., Tsang, P. E., Fang, J., & Zhao, D. (2016). Stabilisation of nanoscale zero-
valent iron with biochar for enhanced transport and in-situ remediation of hexavalent
chromium in soil. Environmental Pollution, 214, 94–100. doi:10.1016/j.envpol.2016.03.072
Su, H., Fang, Z., Tsang, P. E., Zheng, L., Cheng, W., Fang, J., & Zhao, D. (2016).
Remediation of hexavalent chromium contaminated soil by biochar-supported zero-val-
ent iron nanoparticles. Journal of Hazardous Materials, 318, 533–540. doi:10.1016/j.jhaz-
mat.2016.07.039
Sun, F., & Lu, S. (2014). Biochars improve aggregate stability, water retention, and pore-
space properties of clayey soil. Journal of Plant Nutrition and Soil Science, 177(1), 26–33.
doi:10.1002/jpln.201200639
Sun, K., Ro, K., Guo, M., Novak, J., Mashayekhi, H., & Xing, B. (2011). Sorption of bisphe-
nol A, 17a-ethinyl estradiol and phenanthrene on thermally and hydrothermally pro-
duced biochars. Bioresource Technology, 102(10), 5757–5763. doi:10.1016/j.biortech.2011.
03.038
Sun, Y., Gao, B., Yao, Y., Fang, J., Zhang, M., Zhou, Y., … Yang, L. (2014). Effects of feed-
stock type, production method, and pyrolysis temperature on biochar and hydrochar
properties. Chemical Engineering Journal, 240, 574–578. doi:10.1016/j.cej.2013.10.081
Sun, Y., Yu, I. K. M., Tsang, D. C. W., Cao, X., Lin, D., Wang, L., … Li, X. D. (2019).
Multifunctional iron-biochar composites for the removal of potentially toxic elements,
inherent cations, and hetero-chloride from hydraulic fracturing wastewater. Environment
International, 124, 521–532. doi:10.1016/j.envint.2019.01.047
Tan, X., Liu, Y., Zeng, G., Wang, X., Hu, X., Gu, Y., & Yang, Z. (2015). Application of bio-
char for the removal of pollutants from aqueous solutions. Chemosphere, 125, 70–85.
doi:10.1016/j.chemosphere.2014.12.058
Tan, X.-F., Liu, Y.-G., Gu, Y.-L., Xu, Y., Zeng, G.-M., Hu, X.-J., … Li, J. (2016). Biochar-
based nano-composites for the decontamination of wastewater: A review. Bioresource
Technology, 212, 318–333. doi:10.1016/j.biortech.2016.04.093
Tan, Z., Lin, C. S. K., Ji, X., & Rainey, T. J. (2017). Returning biochar to fields: A review.
Applied Soil Ecology, 116, 1–11. doi:10.1016/j.apsoil.2017.03.017
Tang, Q., Shi, C., Shi, W., Huang, X., Ye, Y., Jiang, W., … Li, D. (2019). Preferable phos-
phate removal by nano-La(III) hydroxides modified mesoporous rice husk biochars: Role
of the host pore structure and point of zero charge. Science of the Total Environment,
662, 511–520. doi:10.1016/j.scitotenv.2019.01.159
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY 49

Tariq, F. S., Samsuri, A. W., Karam, D. S., Aris, A. Z., & Jamilu, G. (2019). Bioavailability
and mobility of arsenic, cadmium, and manganese in gold mine tailings amended with
rice husk ash and Fe-coated rice husk ash. Environmental Monitoring and Assessment,
191(4), 232. doi:10.1007/s10661-019-7359-6
Uchimiya, M., Bannon, D. I., & Wartelle, L. H. (2012). Retention of heavy metals by carb-
oxyl functional groups of biochars in small arms range soil. Journal of Agricultural and
Food Chemistry, 60(7), 1798–1809. doi:10.1021/jf2047898
Uchimiya, M., Lima, I. M., Klasson, K. T., & Wartelle, L. H. (2010). Contaminant immobil-
ization and nutrient release by biochar soil amendment: Roles of natural organic matter.
Chemosphere, 80(8), 935–940. doi:10.1016/j.chemosphere.2010.05.020
Uchimiya, M., Lima, I. M., Thomas Klasson, K., Chang, S., Wartelle, L. H., & Rodgers, J. E.
(2010). Immobilization of heavy metal ions (CuII, CdII, NiII, and PbII) by broiler litter-
derived biochars in water and soil. Journal of Agricultural and Food Chemistry, 58(9),
5538–5544. doi:10.1021/jf9044217
Uchimiya, M., Ohno, T., & He, Z. (2013). Pyrolysis temperature-dependent release of dis-
solved organic carbon from plant, manure, and biorefinery wastes. Journal of Analytical
and Applied Pyrolysis, 104, 84–94. doi:10.1016/j.jaap.2013.09.003
Uzoma, K. C., Inoue, M., Andry, H., Zahoor, A., & Nishihara, E. (2011). Influence of bio-
char application on sandy soil hydraulic properties and nutrient retention. Journal of
Food, Agriculture and Environment, 9(3), 1137–1143.
Vijayaraghavan, K. (2016). Biochar: Production strategies, potential feedstocks and applica-
tions. Journal of Environment & Biotechnology Research, 4(1), 41–49.
Wan, Z., Sun, Y., Tsang, D. C. W., Yu, I. K. M., Fan, J., Clark, J. H., … Ok, Y. S. (2019).
A sustainable biochar catalyst synergized with copper heteroatoms and CO2 for singlet
oxygenation and electron transfer routes. Green Chemistry, 21(17), 4800–4814. doi:10.
1039/C9GC01843C
Wang, B., Gao, B., Zimmerman, A. R., Zheng, Y., & Lyu, H. (2018). Novel biochar-impreg-
nated calcium alginate beads with improved water holding and nutrient retention prop-
erties. Journal of Environmental Management, 209, 105–111. doi:10.1016/j.jenvman.2017.
12.041
Wang, D., Guo, W., Zhang, G., Zhou, L., Wang, M., Lu, Y., … Wu, Z. (2017).
Remediation of Cr(VI)-contaminated acid soil using a nanocomposite. ACS Sustainable
Chemistry and Engineering, 5(3), 2243–2254. doi:10.1021/acssuschemeng.6b02569
Wang, J., Liao, Z., Ifthikar, J., Shi, L., Chen, Z., & Chen, Z. (2017). One-step preparation
and application of magnetic sludge-derived biochar on acid orange 7 removal via both
adsorption and persulfate based oxidation. RSC Advances, 7, 18696–18706. doi:10.1039/
C7RA01425B
Wang, L., Chen, L., Tsang, D. C. W., Kua, H. W., Yang, J., Ok, Y. S., … Poon, C. S.
(2019). The roles of biochar as green admixture for sediment-based construction prod-
ucts. Cement and Concrete Composites, 104, 103348. doi:10.1016/j.cemconcomp.2019.
103348
Wang, L., Hou, D., Cao, Y., Ok, Y. S., Tack, F. M. G., Rinklebe, J., & O’Connor, D. (2020).
Remediation of mercury contaminated soil, water, and air: A review of emerging materi-
als and innovative technologies. Environment International, 134, 105281. doi:10.1016/j.
envint.2019.105281
Wang, M. C., Sheng, G. D., & Qiu, Y. P. (2015). A novel manganese-oxide/biochar com-
posite for efficient removal of lead(II) from aqueous solutions. International Journal of
Environmental Science and Technology, 12(5), 1719–1726. doi:10.1007/s13762-014-0538-7
50 S. MANDAL ET AL.

Wang, S., Gao, B., Zimmerman, A. R., Li, Y., Ma, L., Harris, W. G., & Migliaccio, K. W.
(2015). Removal of arsenic by magnetic biochar prepared from pinewood and natural
hematite. Bioresource Technology, 175, 391–395. doi:10.1016/j.biortech.2014.10.104
Wang, S., Zhao, M., Zhou, M., Li, Y. C., Wang, J., Gao, B., … Ok, Y. S. (2019). Biochar-
supported nZVI (nZVI/BC) for contaminant removal from soil and water: A critical
review. Journal of Hazardous Materials, 373, 820–834. doi:10.1016/j.jhazmat.2019.03.080
Wang, X., Gu, Y., Tan, X., Liu, Y., Zhou, Y., Hu, X., … Liu, S. (2019). Functionalized bio-
char/clay composites for reducing the bioavailable fraction of arsenic and cadmium in
river sediment. Environmental Toxicology and Chemistry, 38(10), 2337–2347. doi:10.1002/
etc.4542
Weber, K., & Quicker, P. (2018). Properties of biochar. Fuel, 217, 240–261. doi:10.1016/j.
fuel.2017.12.054
Whitman, T., & Lehmann, J. (2009). Biochar - One way forward for soil carbon in offset
mechanisms in Africa? Environmental Science and Policy, 12(7), 1024–1027. doi:10.1016/
j.envsci.2009.07.013
Wu, C., Cui, M. Q., Xue, S. G., Li, W. C., Huang, L., Jiang, X. X., & Qian, Z. Y. (2018).
Remediation of arsenic-contaminated paddy soil by iron-modified biochar.
Environmental Science and Pollution Research, 25(21), 20792–20801. doi:10.1007/s11356-
018-2268-8
Wu, D., Senbayram, M., Zang, H., Ugurlar, F., Aydemir, S., Br€ uggemann, N., …
Blagodatskaya, E. (2018). Effect of biochar origin and soil pH on greenhouse gas emis-
sions from sandy and clay soils. Applied Soil Ecology, 129, 121–127. doi:10.1016/j.apsoil.
2018.05.009
Wu, H., Lai, C., Zeng, G., Liang, J., Chen, J., Xu, J., … Wan, J. (2017). The interactions of
composting and biochar and their implications for soil amendment and pollution
remediation: A review. Critical Reviews in Biotechnology, 37(6), 754–764. doi:10.1080/
07388551.2016.1232696
Xie, Y., Dong, H., Zeng, G., Tang, L., Jiang, Z., Zhang, C., … Zhang, Y. (2017). The inter-
actions between nanoscale zero-valent iron and microbes in the subsurface environment:
A review. Journal of Hazardous Materials, 321, 390–407. doi:10.1016/j.jhazmat.2016.09.
028
Xiong, X., Yu, I. K. M., Cao, L., Tsang, D. C. W., Zhang, S., & Ok, Y. S. (2017). A review
of biochar-based catalysts for chemical synthesis, biofuel production, and pollution con-
trol. Bioresource Technology, 246, 254–270. doi:10.1016/j.biortech.2017.06.163
Xu, G., Sun, J. N., Shao, H. B., & Chang, S. X. (2014). Biochar had effects on phosphorus
sorption and desorption in three soils with differing acidity. Ecological Engineering, 62,
54–60. doi:10.1016/j.ecoleng.2013.10.027
Xu, X., Zhao, Y., Sima, J., Zhao, L., Masek, O., & Cao, X. (2017). Indispensable role of bio-
char-inherent mineral constituents in its environmental applications: A review.
Bioresource Technology, 241, 887–899. doi:10.1016/j.biortech.2017.06.023
Xu, Y., Fang, Z., & Tsang, E. P. (2016). In situ immobilization of cadmium in soil by stabi-
lized biochar-supported iron phosphate nanoparticles. Environmental Science and
Pollution Research, 23(19), 19164–19172. doi:10.1007/s11356-016-7117-z
Yaashikaa, P. R., Senthil Kumar, P., Varjani, S. J., & Saravanan, A. (2019). Advances in pro-
duction and application of biochar from lignocellulosic feedstocks for remediation of
environmental pollutants. Bioresource Technology, 292, 122030. doi:10.1016/j.biortech.
2019.122030
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY 51

Yan, L., Kong, L., Qu, Z., Li, L., & Shen, G. (2015). Magnetic biochar decorated with ZnS
nanocrytals for Pb (II) removal. ACS Sustainable Chemistry and Engineering, 3(1),
125–132. doi:10.1021/sc500619r
Yang, F., Zhang, S., Sun, Y., Cheng, K., Li, J., & Tsang, D. C. W. (2018). Fabrication and
characterization of hydrophilic corn stalk biochar-supported nanoscale zero-valent iron
composites for efficient metal removal. Bioresource Technology, 265, 490–497. doi:10.
1016/j.biortech.2018.06.029
Yang, F., Zhang, S., Sun, Y., Tsang, D. C. W., Cheng, K., & Ok, Y. S. (2019). Assembling
biochar with various layered double hydroxides for enhancement of phosphorus recov-
ery. Journal of Hazardous Materials, 365, 665–673. doi:10.1016/j.jhazmat.2018.11.047
Yang, X., Yu, I. K. M., Cho, D. W., Chen, S. S., Tsang, D. C. W., Shang, J., … Ok, Y. S.
(2019). Tin-functionalized wood biochar as a sustainable solid catalyst for glucose iso-
merization in biorefinery. ACS Sustainable Chemistry and Engineering, 7(5), 4851–4860.
doi:10.1021/acssuschemeng.8b05311
Yang, Z., Fang, Z., Zheng, L., Cheng, W., Tsang, P. E., Fang, J., & Zhao, D. (2016).
Remediation of lead contaminated soil by biochar-supported nano-hydroxyapatite.
Ecotoxicology and Environmental Safety, 132, 224–230. doi:10.1016/j.ecoenv.2016.06.008
Yao, A., Ju, L., Ling, X., Liu, C., Wei, X., Qiu, H., … Wang, S. (2019). Simultaneous
attenuation of phytoaccumulation of Cd and As in soil treated with inorganic and
organic amendments. Environmental Pollution, 250, 464–474. doi:10.1016/j.envpol.2019.
04.073
Yao, Y., Gao, B., Chen, J., Zhang, M., Inyang, M., Li, Y., … Yang, L. (2013). Engineered
carbon (biochar) prepared by direct pyrolysis of Mg-accumulated tomato tissues:
Characterization and phosphate removal potential. Bioresource Technology, 138, 8–13.
doi:10.1016/j.biortech.2013.03.057
Yao, Y., Gao, B., Inyang, M., Zimmerman, A. R., Cao, X., Pullammanappallil, P., & Yang,
L. (2011a). Biochar derived from anaerobically digested sugar beet tailings:
Characterization and phosphate removal potential. Bioresource Technology, 102(10),
6273–6278. doi:10.1016/j.biortech.2011.03.006
Yao, Y., Gao, B., Inyang, M., Zimmerman, A. R., Cao, X., Pullammanappallil, P., & Yang,
L. (2011b). Removal of phosphate from aqueous solution by biochar derived from anaer-
obically digested sugar beet tailings. Journal of Hazardous Materials, 190(1–3), 501–507.
doi:10.1016/j.jhazmat.2011.03.083
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. doi:10.1016/j.chemosphere.2012.06.002
Ye, S., Yan, M., Tan, X., Liang, J., Zeng, G., Wu, H., … Wang, H. (2019). Facile assembled
biochar-based nanocomposite with improved graphitization for efficient photocatalytic
activity driven by visible light. Applied Catalysis B: Environmental, 250, 78–88. doi:10.
1016/j.apcatb.2019.03.004
Ye, S., Zeng, G., Wu, H., Liang, J., Zhang, C., Dai, J., … Yu, J. (2019). The effects of acti-
vated biochar addition on remediation efficiency of co-composting with contaminated
wetland soil. Resources, Conservation and Recycling, 140, 278–285. doi:10.1016/j.rescon-
rec.2018.10.004
Ye, S., Zeng, G., Wu, H., Zhang, C., Dai, J., Liang, J., … Zhang, C. (2017). Biological tech-
nologies for the remediation of co-contaminated soil. Critical Reviews in Biotechnology,
37(8), 1062–1076. doi:10.1080/07388551.2017.1304357
Ye, S., Zeng, G., Wu, H., Zhang, C., Liang, J., Dai, J., … Cheng, M. (2017). Co-occurrence
and interactions of pollutants, and their impacts on soil remediation—A review. Critical
52 S. MANDAL ET AL.

Reviews in Environmental Science and Technology, 47(16), 1528–1553. doi:10.1080/


10643389.2017.1386951
Yi, Y., Tu, G., Zhao, D., Tsang, P. E., & Fang, Z. (2019). Biomass waste components signifi-
cantly influence the removal of Cr(VI) using magnetic biochar derived from four types
of feedstocks and steel pickling waste liquor. Chemical Engineering Journal, 360,
212–220. doi:10.1016/j.cej.2018.11.205
Yin, D., He, T., Zeng, L., & Chen, J. (2016). Exploration of amendments and agronomic
measures on the remediation of methylmercury-polluted rice in a mercury mining area.
Water, Air, and Soil Pollution, 227, 333. doi:10.1007/s11270-016-3014-9
Yin, D., Wang, X., Peng, B., Tan, C., & Ma, L. Q. (2017). Effect of biochar and Fe-biochar
on Cd and As mobility and transfer in soil-rice system. Chemosphere, 186, 928–937. doi:
10.1016/j.chemosphere.2017.07.126
Yu, I. K. M., Xiong, X., Tsang, D. C. W., Wang, L., Hunt, A. J., Song, H., … Poon, C. S. (2019).
Aluminium-biochar composites as sustainable heterogeneous catalysts for glucose isomerisa-
tion in a biorefinery. Green Chemistry, 21(6), 1267–1281. doi:10.1039/C8GC02466A
Yu, J., Tang, L., Pang, Y., Zeng, G., Wang, J., Deng, Y., … Ren, X. (2019). Magnetic nitro-
gen-doped sludge-derived biochar catalysts for persulfate activation: Internal electron trans-
fer mechanism. Chemical Engineering Journal, 364, 146–159. doi:10.1016/j.cej.2019.01.163
Yu, Z., Qiu, W., Wang, F., Lei, M., Wang, D., & Song, Z. (2017). Effects of manganese
oxide-modified biochar composites on arsenic speciation and accumulation in an indica
rice (Oryza sativa L.) cultivar. Chemosphere, 168, 341–349. doi:10.1016/j.chemosphere.
2016.10.069
Yu, Z., Zhou, L., Huang, Y., Song, Z., & Qiu, W. (2015). Effects of a manganese oxide-
modified biochar composite on adsorption of arsenic in red soil. Journal of
Environmental Management, 163, 155–162. doi:10.1016/j.jenvman.2015.08.020
Yuan, J. H., Xu, R. K., & Zhang, H. (2011). The forms of alkalis in the biochar produced
from crop residues at different temperatures. Bioresource Technology, 102(3), 3488–3497.
doi:10.1016/j.biortech.2010.11.018
Yuan, Y., Bolan, N., Prevoteau, A., Vithanage, M., Biswas, J. K., Ok, Y. S., & Wang, H.
(2017). Applications of biochar in redox-mediated reactions. Bioresource Technology, 246,
271–281. doi:10.1016/j.biortech.2017.06.154
Zama, E. F., Reid, B. J., Arp, H. P. H., Sun, G. X., Yuan, H. Y., & Zhu, Y. G. (2018).
Advances in research on the use of biochar in soil for remediation: A review. Journal of
Soils and Sediments, 18(7), 2433–2450. doi:10.1007/s11368-018-2000-9
Zhang, C., Zeng, G., Huang, D., Lai, C., Chen, M., Cheng, M., … Wang, R. (2019).
Biochar for environmental management: Mitigating greenhouse gas emissions, contamin-
ant treatment, and potential negative impacts. Chemical Engineering Journal, 373,
902–922. doi:10.1016/j.cej.2019.05.139
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. doi:10.1016/j.biombioe.2017.06.024
Zhang, L., Guo, J., Huang, X., Wang, W., Sun, P., Li, Y., & Han, J. (2019). Functionalized
biochar-supported magnetic MnFe2O4 nanocomposite for the removal of Pb(ii) and
Cd(ii). RSC Advances, 9, 365–376. doi:10.1039/C8RA09061K
Zhang, M., & Gao, B. (2013). Removal of arsenic, methylene blue, and phosphate by bio-
char/AlOOH nanocomposite. Chemical Engineering Journal, 226, 286–292. doi:10.1016/j.
cej.2013.04.077
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY 53

Zhang, M., Gao, B., Varnoosfaderani, S., Hebard, A., Yao, Y., & Inyang, M. (2013).
Preparation and characterization of a novel magnetic biochar for arsenic removal.
Bioresource Technology, 130, 457–462. doi:10.1016/j.biortech.2012.11.132
Zhang, N., Fang, Z., & Zhang, R. (2017). Comparison of several amendments for in-site
remediating chromium-contaminated farmland soil. Water, Air, and Soil Pollution, 228,
400. doi:10.1007/s11270-017-3571-6
Zhang, R., Zhang, N., & Fang, Z. (2018). In situ remediation of hexavalent chromium con-
taminated soil by CMC-stabilized nanoscale zero-valent iron composited with biochar.
Water Science and Technology, 77(5-6), 1622–1631. doi:10.2166/wst.2018.039
Zhang, X., Wang, H., He, L., Lu, K., Sarmah, A., Li, J., … Huang, H. (2013). Using bio-
char for remediation of soils contaminated with heavy metals and organic pollutants.
Environmental Science and Pollution Research, 20(12), 8472–8483. doi:10.1007/s11356-
013-1659-0
Zhao, B., O’Connor, D., Zhang, J., Peng, T., Shen, Z., Tsang, D. C. W., & Hou, D. (2018).
Effect of pyrolysis temperature, heating rate, and residence time on rapeseed stem
derived biochar. Journal of Cleaner Production, 174, 977–987. doi:10.1016/j.jclepro.2017.
11.013
Zhao, J. J., Shen, X. J., Domene, X., Alca~
niz, 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, 9869. doi:10.1038/s41598-019-
46234-4
Zheng, J., Chen, J., Pan, G., Liu, X., Zhang, X., Li, L., … Jinwei, Z. (2016). Biochar
decreased microbial metabolic quotient and shifted community composition four years
after a single incorporation in a slightly acid rice paddy from southwest China. Science
of the Total Environment, 571, 206–217. doi:10.1016/j.scitotenv.2016.07.135
Zhou, Y., Gao, B., Zimmerman, A. R., Fang, J., Sun, Y., & Cao, X. (2013). Sorption of
heavy metals on chitosan-modified biochars and its biological effects. Chemical
Engineering Journal, 231, 512–518. doi:10.1016/j.cej.2013.07.036
Zhou, Y., Xiang, Y., He, Y., Yang, Y., Zhang, J., Luo, L., … Tang, L. (2018). Applications
and factors influencing of the persulfate-based advanced oxidation processes for the
remediation of groundwater and soil contaminated with organic compounds. Journal of
Hazardous Materials, 359, 396–407. doi:10.1016/j.jhazmat.2018.07.083
Zhu, S., Huang, X., Ma, F., Wang, L., Duan, X., & Wang, S. (2018). Catalytic removal of
aqueous contaminants on N-doped graphitic biochars: Inherent roles of adsorption and
nonradical mechanisms. Environmental Science and Technology, 52, 8649–8658. doi:10.
1021/acs.est.8b01817
Zhu, X., Chen, B., Zhu, L., & Xing, B. (2017). Effects and mechanisms of biochar-microbe
interactions in soil improvement and pollution remediation: A review. Environmental
Pollution, 227, 98–115. doi:10.1016/j.envpol.2017.04.032
Zhuang, L., Li, Q., Chen, J., Ma, B., & Chen, S. (2014). Carbothermal preparation of porous
carbon-encapsulated iron composite for the removal of trace hexavalent chromium.
Chemical Engineering Journal, 253, 24–33. doi:10.1016/j.cej.2014.05.038
Zimmerman, A. R., Gao, B., & Ahn, M. Y. (2011). Positive and negative carbon mineraliza-
tion priming effects among a variety of biochar-amended soils. Soil Biology and
Biochemistry, 43(6), 1169–1179. doi:10.1016/j.soilbio.2011.02.005
Zou, Y., Wang, X., Khan, A., Wang, P., Liu, Y., Alsaedi, A., … Wang, X. (2016).
Environmental remediation and application of nanoscale zero-valent iron and its compo-
sites for the removal of heavy metal ions: A review. Environmental Science and
Technology, 50, 7290–7304. doi:10.1021/acs.est.6b01897

You might also like