Download as pdf or txt
Download as pdf or txt
You are on page 1of 43

Critical Reviews in Environmental Science and

Technology

ISSN: (Print) (Online) Journal homepage: https://www.tandfonline.com/loi/best20

Enrichment of primary macronutrients in biochar


for sustainable agriculture: A review

Adnan Asad Karim, Manish Kumar, Ekta Singh, Aman Kumar, Sunil Kumar,
Arati Ray & Nabin Kumar Dhal

To cite this article: Adnan Asad Karim, Manish Kumar, Ekta Singh, Aman Kumar, Sunil Kumar,
Arati Ray & Nabin Kumar Dhal (2021): Enrichment of primary macronutrients in biochar for
sustainable agriculture: A review, Critical Reviews in Environmental Science and Technology, DOI:
10.1080/10643389.2020.1859271

To link to this article: https://doi.org/10.1080/10643389.2020.1859271

View supplementary material

Published online: 06 Jan 2021.

Submit your article to this journal

Article views: 281

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/10.1080/10643389.2020.1859271

Enrichment of primary macronutrients in biochar for


sustainable agriculture: A review
Adnan Asad Karima, Manish Kumara, Ekta Singhb, Aman Kumarb,
Sunil Kumarb, Arati Raya, and Nabin Kumar Dhala
a
Environment and Sustainability Department, CSIR-Institute of Minerals and Materials
Technology, Bhubaneswar, India; bTechnology Development Centre, CSIR-National Environmental
Engineering Research Institute (CSIR-NEERI), Nagpur, India

ABSTRACT
Macronutrient-enriched biochars
have potential for nutrient recy-
cling from waste, enhance soil
fertility, reduce consumption of
fertilizers, and thereby promoting
sustainable agriculture. The focus
of the recent research studies has
been on exploring different
methods for production of mac-
ronutrients (nitrogen [N], phos-
phorus [P], and potassium [K])
enriched biochar fertilizers. There
is now an urgent need to comprehend the work carried out on biochar fertilizers for a
proper scientific understanding and advancement. This review provides crucial informa-
tion about nutrients enriched biochars production and the underlying mechanisms
toward its use in agriculture. The production methods, nutrient characteristics, potential
major challenges and future perspectives of using these nutrients enriched biochars
have also been dealt in the present review. Among primary macronutrients, the main
focus of nutrients enriched biochar are on N and P. Total N, P, and K enriched up to
10%, 15.32%, 25.85% in biochar through controlled (primarily temperature) pyrolysis
and utilization of specific feedstock. Although total macronutrients content in nutrients
enriched biochar were thoroughly studied, but the bio-labile and plant available frac-
tions of these nutrients are yet to be extensively researched. Nutrients transformation
in biochars revealed about comparatively higher bio-available K fraction due to forma-
tion of water soluble kalicinite (KHCO3) minerals. The efficacy of biochars toward soil
fertility and plant growth was widely studied (mainly as pot experiments) and there
results varied. Overall, the prospects of nutrients enriched biochars toward sustainable
agriculture has gained momentum with time and further engineered biochars hold
great potential to cater for specific soil-plant-environment system.

KEYWORDS Agriculture; biochar fertilizers; macronutrients; soil fertility; waste

CONTACT Manish Kumar manish@immt.res.in; Adnan Asad Karim md.karime@gmail.com Environment


and Sustainability Department, CSIR-Institute of Minerals and Materials Technology, Bhubaneswar, Odisha
751013, India.
Adnan Asad Karim and Manish Kumar are the joint first authors.
Supplemental data for this article can be accessed at publisher’s website.
ß 2021 Taylor & Francis Group, LLC
2 A. A. KARIM ET AL.

1. Introduction
Sustainable agriculture seeks development and implementation of different
agricultural management strategies to alleviate detrimental effects of inten-
sified land use. In recent years, extensive research focused on efficient util-
ization of various natural and synthetic fertilizers for the sustainable
agriculture. These fertilizers are extensively utilized to enhance the soil fer-
tility and crop productivity (Semida et al., 2019). Synthetic fertilizers con-
tains nutrients (nitrogen-N, phosphorus-P, potassium-K) in high quantity
and water soluble forms, therefore widely used in soils compared to man-
ures, and organic fertilizers (Olsen, 1978; Vanlauwe et al., 2010). However,
continuous and overuse of synthetic fertilizers (specifically superphosphate)
resulted in accumulation of toxic heavy metals (Cu, Zn, Cd, Pb, As, and
Mn) in soils and crops, posing health risks to humans (Gimeno-Garcia
et al., 1996; Kelepertzis, 2014; Lin et al. 2019; Modaibsb & Al-Sewailem,
1999). Therefore, intensive farming practices with sole dependency on con-
ventional fertilizers are generally unfavorable and costly for soil quality and
ecosystem (Ding et al., 2016; Karer et al., 2015). The use of eco-friendly
approach to improve fertility of the soil is therefore imperative (Inyang
et al., 2016; Ok et al., 2015).
Biochars application in soil received growing interest over past two deca-
des for sustainable agriculture by improving fertility, productivity (Peiris
et al. 2019) along with numerous advantages like waste management, and
climate change mitigation (Awad et al., 2018; El-Naggar et al., 2019).
Biochars are carbonaceous solid material produced through thermo-chem-
ical conversion of organic substances in presence of limited oxygen for soil
application (Kumar, Singh, Singh, et al., 2020; Lehmann & Joseph, 2015).
Application of biochar are widely studied, but its efficacy is not always
positive and showed variation depending on characteristics along with soil-
environmental conditions (El-Naggar et al., 2019; Kumar, Singh, Khapre,
et al., 2020; Purakayastha et al., 2019; Singh et al., 2020). Characteristic of
biochars are mainly influenced by chemical composition of feedstock and
production conditions. Biochars produced from hardwood wastes like euca-
lyptus, pine bark and sawdust possess higher alkaline pH, carbon (C) con-
tent, and aromaticity; whereas, ash content and nutrients like N, P, K, and
calcium (Ca) are lesser (Domingues et al., 2017). Biochars produced from
crop residues (Prakongkep et al., 2015; Wang, Hu, et al., 2013) and animal
manures (Cantrell et al., 2012) were contrastingly alkaline with higher
nutrients, but with lesser C content and aromaticity. In relation to pH
value, acidic and neutral biochars were also reported (Chan & Xu, 2012).
Low temperature (250–350  C) biochars of Birch (Betula spp.) (Hagner
et al., 2016), Pecan shell and Switchgrass (Novak et al., 2009) bear acidic
pH (5–6), whereas Eucalyptus deglupta (Rondon et al., 2007) derived
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY 3

biochar showed neutral pH (7.0–7.5). When pyrolysis temperature


increased from 300 to 800  C, C content in biochars increased in contrast
to hydrogen and oxygen (Kookana et al., 2011). The porous structure of
biochars augment beneficial soil microorganism (e.g., mycorrhizae and bac-
teria) population due to their protective habitat (Atkinson et al., 2010;
Spokas et al., 2012) which helps in nutrients transformation and dynamics.
Biochars having less C and high ash content improved plant productivity
in sandy and acidic soils due to biochar induced liming effect, improved
soil physical structure and nutrient use efficiency (Dai et al., 2020).
Biochars derived from less C/N ratio feedstocks (like animal manure) were
more efficient in improving P availability in soils (Gao et al., 2019). Earlier
studies revealed about combined application of fertilizers (inorganic/organic
amendments) with biochar lead to improve crop yield (Ye et al., 2020). As
an alternative to fertilizers, several researchers focused on nutrients rich
biochar production and application for improving soil fertility and product-
ivity (Chan & Xu, 2012; Ippolito et al., 2015). Application of nutrients rich
biochars could be more useful for degraded and/or macronutrients def-
icit soils.
To better comprehend the ongoing research on nutrients rich biochar
fertilizer, relevant literatures were reviewed to gauge research progress and
to identify the knowledge gaps for highlighting future research directions.
Earlier published reviews (El-Naggar et al., 2018, 2019) on biochars
included summary of its nutrients value with emphasis for potential fertil-
izer application. However, a review exclusively on the primary macronu-
trients (NPK) enrichment in biochars not yet reported properly and in
view of that covered as overarching theme of the current review article.
The specific goals of present review article are as follows:

1. Co-relation between the various production processes and enrichment


of primary macronutrients in biochars to determine suitable feedstocks,
process conditions, limitations and gaps
2. Mechanisms responsible for enrichment and transformation of primary
macronutrients forms in biochars is outlined and specified to under-
stand the nutrients release dynamics in soils
3. Specifying the contrasting research findings, shortcomings and associ-
ated negative outcomes based on diversified characteristics and perform-
ance biochars in soils
4. Discussion about biochars production and application in view of farm-
ers’ socio-economic aspects and scenario
5. Environmental concerns associated with the biochars application in soil
like potentially toxic constituents etc.
6. Future research directions regarding the best possible ways to produce
different types of nutrients enriched improved biochars as fertilizer
4 A. A. KARIM ET AL.

2. Production of nutrients enriched biochar-based fertilizer and


underlying transformation mechanisms
2.1. Methods for nutrients enriched biochar-based fertilizer production
Biochars fertilizer contains plant nutrients with potential to improve soil
fertility for better plant growth, development and productivty. Different
methods were documented regarding production of nutrients rich biochar
fertilizer. These methods (Figure 1) broadly classified into three categories:
(1) direct treatment, (2) pre-treatment, and (3) post-treatment.

2.1.1. Direct treatment method


The direct treatment method involves direct thermo-chemical conversion
of nutrient rich feedstocks into biochar fertilizer. Table S1 present charac-
teristics of biochar fertilizers produced from different nutrient rich feed-
stocks, which has potential for soil application. Almost all research studies
reported that, feedstocks inherently rich in nutrients undergo slow pyrolysis
leading to nutrients enrichment in biochar. Pyrolysis temperature were
dominant factor for nutrient enrichment in biochars due to varied volatil-
ization temperature of nutrients. For N enrichment in biochar, relatively
lower temperature range 300–400  C were more appropriate, whereas ele-
vated temperature of around 700  C was suitable for P and K (Biederman
& Harpole, 2013; Chan & Xu, 2012). Several biomass feedstocks such as
poultry litter, swine manure, anaerobic digestate, bone-meal, algae were
explored to produce nutrient rich biochars. In subsequent paragraphs,

Figure 1. Schema of different methods for production of biochars fertilizer.


CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY 5

pyrolytic methods to produce primarily N-P-K, P, and K enriched biochars


are discussed in detail.
Production of macronutrients rich biochars from human manure
(300–700  C), swine manure (400–800  C), poultry manure/litter
(400–600  C) constituted (in %) total N (2.4–4.8, 1.6–3.2, and 2.0–2.8), P
(5.4–8.1, 6.1–7.7, and 4.0–5.8) and K (1.9–2.6, 2.7–3.1, and 3.8–5.8) (Liu
et al., 2014; Subedi, Taupe, Ikoyi, et al., 2016; Tsai et al., 2012). Chlorella-
based microalgal residue (generated from cell disruption and dehydration
process) biochar contained remarkable high N (>10%) at 400  C and its
value decreased with increase in temperature (Chang et al., 2015). Lou et al.
(2017) studied spent mushroom substrate compost derived biochars rich in
total calcium (15.17–17.76%), K (3.16–4.45%), N (4.1–4.7%), and P
(1.19–3.42%). In contrast to above mentioned N, P and K containing bio-
chars, wastewater sludge biochars retained 1.2–3.3% N along with
4.47–6.17% sulfur (S), and 3.47–5.35% Ca (Hossain et al., 2011).
Bio-fermentation waste (Escherichia coli bacterial biomass) produced P
rich (8.4%) biochars at 600  C, which also contained about 8.6% N (Kim
et al., 2018). Zwetsloot et al. (2016) revealed production of P rich biochars
from bone waste at 350  C (12.71%) and 750  C (15.32%). Ma and
Matsunaka (2013) study about dairy cattle carcasses (mixture of skin, meat,
and bone) biochars (450  C) showed total P 10.8%. Yuan et al. (2015) also
highlighted about higher P in biochars produced from sewage sludge at
higher pyrolysis temperature (at 300  C-3.88% and at 700  C-4.92%),
whereas N showed reverse trend (6.1% at 300  C and 0.9% at 700  C).
K enriched biochars production from banana peduncle biomass waste
was reported by Karim et al. (2017). In place of widely explored slow pyr-
olysis method, thermal plasma fast processing with regulated temperature
varying processing time (3–9 min) were developed. Maximum total K con-
tent 25.85% in biochar was achieved at 7 min under argon plasma pyrolysis.
Roberts et al. (2015) reported about K enriched biochars from six different
species of seaweeds (brown seaweeds: Saccharina, Undaria and Sargassum;
red seaweeds: Gracilaria, Kappaphycus and Eucheuma) produced at 450  C
pyrolysis temperature. Biochars derived from red seaweeds contained rela-
tively higher total K (5.12–16.3%) than the brown seaweeds that also
retains total P up to 5.19%.
The average values of total NPK contents in biochars fertilizer of differ-
ent feedstocks are shown in Figure 2. Among different feedstocks, algal
waste derived biochars contains relatively very high total N with relatively
lower total C/N ratio of 3.79–8.78 (Zhou et al., 2019). However, organic C/
N were not extensively reported for biochars, which could have indicated
actual potential bio-availability of N in soils. Bone waste was found to be a
more suitable feedstock for the production of P rich biochars. Banana
6 A. A. KARIM ET AL.

Figure 2. Average values (calculated from Table S1 data) of total N, P, and K content (as elem-
ental %) in biochar fertilizer derived from various feedstocks.

peduncle was the better feedstock for K rich biochars production. Biochars
derived from bio-fermentation waste were rich in both N and P nutrients.
Cattle carcasses could be suitable feedstock for the production of biochar
rich in multiple macronutrients (NPK). However, a major knowledge gap
is total bio-available fraction of NPK in biochars are rarely studied and not
known properly. Few studies investigated about P and reported less water
soluble fraction compared to total content in biochars, which could be due
to its fixed chemical form and trapping in the particles (Song & Guo,
2012). Therefore, it is crucial that both total content and bio-available frac-
tion should be reported by researchers for nutrient rich biochars to identify
its actual potential for soil application.

2.1.2. Pre-treatment method


In pre-treatment method, the biomass is treated with nutrient rich materi-
als (e.g., nutrient rich minerals, chemical fertilizers and fertilizer industry
waste, animal waste) prior to thermo-chemical processing (high pyrolysis
temperature about 600  C) for production of biochar fertilizers (Joseph
et al., 2013). Research studies on pretreatment method are very limited.
Zhao et al. (2016) evaluated the influence of mixing phosphate fertilizers
(bone meal and triple super phosphate) with sawdust and switch grass bio-
mass on characteristics of biochars produced through slow pyrolysis
(600  C). The biochars contained high amount of P along with high C
retention and enhanced heavy metal stabilization in soils. During pyrolysis,
Ca(H2PO4)2 (present in triple super phosphate) transformed to Ca2P2O7,
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY 7

which reacts with biomass C to form C–P and C–O–PO3. These reactions
helped in improving the C and P retention in biochars, and its application
as slow P release fertilizer. Mixture of banana peduncle biomass and phos-
phogypsum was treated through thermal plasma processing for 7 min for
production of K-S rich biochars. The biochars contained high contents of
K (4.2–12.7%) and S (13.3–17.8%), along with reduced bioavailable frac-
tions of toxic contaminants like fluoride, cadmium, lead etc. K and S were
mainly present as potassium sulfate (highly water soluble) in biochars,
which indicates the fast release of K and S from biochars for enhanced
plant uptake in soils (Karim, Kumar, Mohapatra, & Singh, 2019). Biochar
fertilizers were produced through thermal treatment of K rich banana ped-
uncle biomass and P rich effluent sludge mixture (Karim, Kumar,
Mohapatra, Singh, & Panda, 2019). Biochars produced through plasma
processing (1200–1500  C) contained more total P (11.81%) and K (21.9%)
contents in comparison to slow pyrolysis performed at 700  C (4.2% P,
11.5% K). It was due to the comparatively elevated temperature during
plasma processing. Bioavailable P fractions were 0.6 and 0.56% in biochars
produced via plasma processing and slow pyrolysis respectively. Overall, all
above studies demonstrated that pre-treatment method could be a suitable
way for the production of designer biochar fertilizers such as biochars
enriched in multiple macro/micro nutrients for application in specific soil-
plant systems.

2.1.3. Post-treatment method


In post-treatment method, biochars were treated with a nutrient rich source
(e.g., wastewater, waste gases) at ambient conditions or temperature regu-
lated conditions to prepare nutrient enriched biochar fertilizers (Joseph
et al., 2013). Research studies have been focused on the utilization of bio-
chars as the low-cost sorbent for nutrient recovery from wastewater
(Antunes et al., 2017; Gai et al., 2014; Kizito et al., 2015; Shepherd et al.,
2017; Zhang & Wang, 2016). Post saturation, the nutrient loaded biochar
were proposed to be used as fertilizers in soils. It can help in reducing
nutrient pollution and eutrophication along with promoting
nutrients cycling.
Biochars produced (300–600  C) from various feedstocks were evaluated
for nutrient (Phosphate and Ammonium) sorption from liquid wastes and
presented in Table S2. Diverse feedstocks used for biochars were anaerobic
digestate sewage sludge, bamboo, corn waste, cotton stalks waste, grapevine
cane, greenhouse waste, hardwood shavings, municipal waste, oak wood,
saw dust, rice husk, corn straw, wheat straw, peanut shells, water treatment
sludge, sugar beet tailings, wood waste, and Thalia dealbata. The liquid
wastes used as nutrient rich sources for sorption were included urine, swine
8 A. A. KARIM ET AL.

manure anaerobic digestate liquid/slurry, biogas fermentation liquid, and


wastewater. Several researchers also evaluated pyrolysis of feedstock/biomass
pretreated with Magnesium (Mg) solutions to produce Mg rich biochars for
enhancement of its phosphate adsorption capacity (Cui et al., 2016; Fang
et al., 2015; Yao et al., 2011a, 2013). Among biochars, Hardwood shavings
biochar (300  C) exhibited lowest sorption capacity both for ammonium
(5.3 mg/g) and phosphate (0.24 mg/g) (Sarkhot et al., 2013). Municipal waste
biochar showed relatively highest ammonium sorption capacity (137.3 mg/g)
(Takaya et al., 2016). Biochar produced at pyrolysis temperature of 450  C
was more suitable for higher ammonium sorption capacity, further increase
in pyrolysis temperature to 650  C decreased its sorption capacity.
Ammonium adsorption capacity of biochars was found to be highly depend-
ent on the reactions with oxygen functional groups, CEC, compared to sur-
face adsorption (physisorption) (Mukherjee et al., 2011; Takaya et al., 2016).
During sorption process, MgO and P2O5 present in biochars are released as
Mg2þ and PO43 into solutions, which combines (at pH range 7–11) with
NH4þ to form Struvite (MgNH4PO4) precipitation (Equations (1) and (2)).
MgO þ 2NH4 þ þ H2 O ! Mg2þ þ 2NH3 :H2 O (1)
Mg2þ þ NH4 þ þ PO3
4 þ 6H2 O ! MgNH4 PO4 :6H2 O (2)
The availability of more polar functional groups (ACOOH, AOH, CAO)
in low temperature (300  C) biochars is responsible for urea sorption (from
human urine) capacity, whereas high temperature biochars (500–600  C)
showed better adsorption capacity, if urea concentration gradient is high
(above 4000 mg/L in urine) due to diffusion of urea in the biochars with
high surface area. The plant germination experiments further indicated that
exhausted biochar could be used for improving soil productivity (Zhang
et al., 2015). In a study (Zhang et al., 2017), S rich biochar (Sulfachar) was
prepared by using biochar for capturing toxic hydrogen sulfide gas from
biogas emissions stream, which generated during anaerobic digestion of
organic waste. The Sulfachar collected large amount (36.5%) of S in both
elemental and sulfate form. In addition, adsorbed S was in plant available
form and Sulfachar treated soils showed increase in the uptake of S and
other nutrients in corn plant.
Non woody biomass and high pyrolysis temperature (600–800  C) were
found more favorable showing high adsorption capacity for phosphate.
Corn waste biochars have higher (225 mg/g) phosphate sorption capacity
(Fang et al., 2014). After modifying corncob biochars by enriching Ca-Mg,
the sorption capacity increased up to 326.62 mg/g (Fang et al., 2015).
Montmorillonite (Chen et al., 2017) and ocher (Shepherd et al., 2016) min-
erals were used with biomass to produce biochar composites to provide
various surface-active sites for effective adsorption. Adsorption of
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY 9

phosphate mainly occurred through electrostatic attraction and bonding


with metal cations like Ca2þ, Mg2þ, and Fe3þ present in biochars (Chen
et al. 2017; Fang et al., 2014, 2015; Takaya et al., 2016; Veni et al., 2017)
and ion exchange with hydroxyl groups oxygen rich functional groups (car-
bonyl, carboxyl, hydroxyl, and methoxyl) present in biochar surfaces
(Sarkhot et al., 2013). Precipitation and retention of phosphate as brushite
(CaHPO3) mineral form also identified in biochars (Marshall et al., 2017).
The presence of Mg nanocrystals and/or MgO nanoparticles on biochar’s
surface could further enhanced its phosphate sorption capacity by provid-
ing more active sites and P precipitation as magnesium phosphates (Yao
et al., 2011a, 2013). Presence of Ca in biochar does not have a prominent
effect on phosphate sorption through precipitation as calcium phosphate. It
could be due to the calcium existence as less water-soluble calcite mineral
and entrapment inside the biochar structures, which limited the release of
Ca from biochars into solution due to chemical precipitation (Cao &
Harris, 2010; Yao et al., 2011b). Desorption experiment revealed 97% of
phosphate recovery from biochars mostly as calcium phosphate by modify-
ing solution pH 4 (Marshall et al., 2017). Phosphate effectively desorbed
from biochar in both acidic (75–88%) and neutral pH (57–78%), which
indicated reversible sorption process (Kizito et al. 2017). In pot culture
experiments, the post-sorption biochar exhibited improved P content in
soil and plant (Marshall et al., 2017). P was present in bioavailable form in
biochars. Due to this, post-sorption biochar also improved the germination
and growth of grass seedlings (Yao et al., 2013).
Nutrient binding capabilities of the specific biochar type are highly
influenced by its properties, including pH, surface area and ion-exchange
capacities (Ding et al., 2016). The mechanisms for biochars binding cap-
acity of polar and non-polar compounds are mainly due to hydrophobic
bonding, a-a electrons donor interactions as a result of fused structures
and the weak H bonding (Conte et al., 2013). The process of sorption
and desorption of nutrients in biochars controls its release in soil solu-
tion and bioavailability. The nutrient (especially adsorbed nutrients)
release from biochar through desorption process is heavily affected by
soil type, feedstocks, process of formation and biochar application rates.
The soil pH, organic matter, and activity of cations (Al3þ, Fe3þ and
Ca2þ) that interact with nutrients (phosphate and nitrate) in biochar
also influences their bioavailability in soils (Xu et al., 2014). The phys-
ical characteristics, organo-mineral complexes and various interactions
between biochar and soil also have a strong impact on the dynamic of
release and maintenance of nutrients. Nutrient binding and its release
are mainly due to the following mechanisms (Schmidt et al., 2017; Xiao
& Pignatello, 2016):
10 A. A. KARIM ET AL.

1. Ionic bonding i.e. direct positive ion bonding, micro sites negative
charge, surface functional groups or precipitated metal or pyrolytic
oxides tars
2. Water bridges between dissolved ion hydraulic shells and systems, like
oxygen-generated organic groups on biochar surfaces
3. Covalent bonding/chemical bonding involves sharing of electron pairs
among ions on the biochar
4. Physical entrapment (minerals entering the pore spaces)
5. Intra-molecular hydrogen bonds between the electron rich proton
acceptors and donors on the surface of biochar
6. Van-der-Waals interactions between the hydrophobic characteristics of
an organic molecules and the organic poly-aromatic components of
the biochar

2.2. Nutrients transformation mechanisms in biochar fertilizer production


During pyrolytic decomposition of biomass constituents, nutrient enrich-
ment in biochars mainly takes place because of the formation of thermally
resistant form and volatilization of other compounds (Domingues et al.,
2017). Minerals form of nutrients in biochars mainly determined its avail-
ability to plants. In addition, the occurrence of nutrients on the surface
and inside the biochar particles can also influence its plant availability
(Keiluweit et al., 2010). In previous studies, researchers have mainly studied
the nutrients (NPK) transformation during biochar production through dir-
ect treatment method, explained in subsequent sub-sections. Studies related
to nutrients transformation in biochar produced through pre-treatment and
post-treatment methods are scarce.

2.2.1. N transformation
Several studies had documented the mechanisms for transformation of N
species from biomass to biochars, which was influenced by biomass com-
position, pyrolysis atmosphere, and temperature. Loss of N from biomass
feedstock mainly initiated at pyrolysis temperature of 150  C, due to the
degradation of hemicellulose and cellulose fractions. The degradation fur-
ther intensified gradually with the rise in temperature from 250 to 500  C,
which causes major loss of nutrients due to volatilization (Tian et al.,
2017). Transformation of N due to the pyrolysis and its dominant forms
present in the biochar fertilizers is shown in Figure 3. Tian et al. (2014)
examined the N speciation in sewage sludge biochars produced at
150–800  C through pyrolysis. Raw sludge mainly consisted of proteins,
pyridinic, pyrrolic, and amine forms of N. The pyrolysis of sludge pro-
duced biochar mainly rich in heterocyclic-N derived from pyrrolic and
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY 11

Figure 3. N species transformation due to pyrolysis and its dominant forms present in the bio-
char fertilizers.

pyridinic N of raw sludge. In addition, the sludge biochars contained


amine, nitrile, and other organic N compounds. Transformation of N
forms also studied from rice straw and its biochars (Liu, Tan, Zhang, et al.,
2018). Rice straw found to be dominantly composed of proteins (70%) and
remaining N are present as alkaloids, amines, and ammonium. In its bio-
char, pyrolysis process changed the N species to organic (25–44% pyrrole,
12–23% pyridine, 4–12% amino, and below 10% nitrile), and inorganic
forms (NH4þ, NO3, and NO2). Liu, Tan, and Ye (2018) investigated the
N speciation and transformation in wheat straw biochars produced at dif-
ferent pyrolysis temperature ranged 300–800  C using isotope tracer and
XPS analysis. This study concluded that the wheat straw and its biochars
produced at temperature range 300–500  C contained similar N compounds
i.e., alkaloid, free amino acid, protein, and NH4þ. However, gradual trans-
formation of these N species formed additional N compounds mainly
amino, pyridine, pyrrole, quaternary N, NH4þ, NO2, and NO3 in bio-
chars. A study by Tian et al. (2017) concluded retention of N is relatively
higher in the CO2 atmosphere than N2 because it can inhibit the formation
of volatile species such as NO, NH3, HCN compounds etc. Thus, the CO2
atmosphere and lower pyrolysis temperature (Preferably around 300  C)
generally favored the preparation of N enriched biochar.

2.2.2. P transformation
During pyrolysis of biomass, organic P (phosphate monoesters, mostly
phytic acid) compounds transformed into inorganic P oxides and metal
(Ca, Mg, Fe) phosphates (Dai et al., 2016). The volatilization of P mostly
started above 250  C, thereafter increase in pyrolysis temperature caused
12 A. A. KARIM ET AL.

changes in the proportion of organic, inorganic, and volatile P compounds.


The enriched P fraction in biochar mainly constitutes of inorganic P frac-
tion due to the thermal transformation of organic P into inorganic com-
pounds. Conversion of organic P species into inorganic species is
comparatively higher in N2 presence than CO2. Above 700  C pyrolysis
temperature, inorganic P mostly transformed into volatile species leading to
its loss from biochar (Tian et al., 2017).
Several researchers investigated the speciation and transformation of P
during pyrolysis in sewage sludge biochars. Huang and Tang (2015)
observed the raw sludge predominantly contained orthophosphate and
long-chained poly-phosphate. The pyrolysis transformed these P species
into relatively less extractable forms i.e. pyro-phosphate and short-chained
poly-phosphate in biochar. The extractability of P substantially reduced
with increased pyrolysis temperature (250–600  C) to produce biochars.
Another study by Qian and Jiang (2014) explained the impact of fast pyr-
olysis temperature range (400–800  C) and gaseous conditions (N2, CO2,
and air) on P content in sewage sludge biochar. They found that pyrolysis
temperature predominantly affects the formation of P-species in biochars
and not the gaseous atmosphere. Moreover, they elucidated (based on XRD
and solid-state 31P NMR) the pathway for formation of varied P species in
sewage sludge biochars. P mainly presents as organic and inorganic P
(orthophosphate and pyro-phosphate) in sewage sludge. However, among
them, orthophosphate constitutes major proportion (75%) of sewage sludge.
The formation of pyrophosphates in biochars is dominant around
400–600  C pyrolysis temperature due to dehydration of M2(HPO4)x (or
M(H2PO4)x) present in sludge. In contrast, pyro-phosphates formation
largely declined and metal (Ca, Mg) phosphates or P bearing minerals
(Stanfieldite-Ca4(Mg, Fe)5(PO4)6) dominated at elevated temperature
around 700–800  C. This study used sequential extraction process that
includes NaHCO3, NaOH and HCl extractants to predict the availability of
different P forms. The soluble orthophosphate, pyro-phosphate and weakly
bound forms of P extracted in the solution using NaHCO3. Adsorbed phos-
phate anions on positively charge metals Al/Fe and some insoluble metal
phosphate compounds extracted in the NaOH solution. Whereas, insoluble
Ca/Mg phosphate compounds, P bearing minerals and P trapped in the lat-
tice of minerals mostly extracted using HCl solution. As the pyrolysis tem-
perature rises, the water-soluble P transformed into NaHCO3-P, NaOH-P
and HCl-P with decreasing solubility. P speciation in biochars produced
from two different types of sewage sludge samples under varied thermal
processing conditions was investigated (Steckenmesser et al., 2017). First
sample was sewage sludge formed through precipitation method (using
iron chloride) for treatment of industrial wastewater. Second sample is the
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY 13

sewage sludge formed by use of biological precipitation method for rural


wastewater treatment and post-stabilization using aluminum sulfate. Both
sewage sludge samples were thermally treated at low temperature ranged
400–500  C. Mineralogical analysis (using XRD) of biological sludge not
exhibited any presence of crystalline P compounds, which could be due to
organic polyphosphates and amorphous P compounds as major constitu-
ents. Vivianite (Fe3(PO4)2.8H2O) and lipscombite (Fe3(PO4)2(OH)2) miner-
als were evident in industrial sewage sludge. The evaluation of total P
fractionation showed that the chemical sludge contains 41.5% Fe-P (NaOH
extract), 20% Al-P (NH4F extract), 17% Ca-P (H2SO4 extract), and 2.5%
water-soluble extract. On the other hand, biological sludge constitutes of
47.5% Al-P, 17% Fe-P, 11% Ca-P, and 3.3% water soluble P.
The P transformation in poultry litter biochars produced through slow
pyrolysis (300–600  C) was researched by Li et al. (2018). Raw poultry litter
comprised of organic (phospholipids, phytates, and nucleic acids) and inor-
ganic P species i.e. CaHPO4, Ca3(PO4)2, and Ca8H2(PO4)65H2O. Pyrolysis
specifically above 400  C drastically declined the organic forms of P in bio-
chars due to its decomposition and subsequent transformation into inor-
ganic forms. Initially, organic P (water-soluble and NaOH extractable
fraction) converted to inorganic pyro-phosphates and other poly-phos-
phates, which further at higher pyrolysis temperature changed dominantly
to orthophosphates. The poly-phosphates to orthophosphates conversion
were facilitated by availability of metal ions at higher pyrolysis temperature.
The low labile and water-insoluble hydroxyapatite and oxyapatite minerals
(HCl extractable) formed in biochars. In addition, surface precipitates and
complexes of P on calcium carbonate and Fe/Al oxides (NaHCO3 and
NaOH extractable), which are moderately labile are marginally present in
biochars. These changes in P speciation due to pyrolysis transformed the
readily labile and water-soluble P fraction in poultry litter to less labile and
mobile fractions in biochars, which could be beneficial to restrict its loss
through leaching and run off, and thus enhancing its plant availability.
Transformation of P in biochars derived from peanut husk, wheat and
maize straw through slow pyrolysis (300–600  C) was studied by Xu,
Zhang, Shao, and Sun (2016). Water-soluble P forms are dominant in bio-
chars produced at low temperature (up to 400  C), which are transformed
into more labile form (400 to 600  C) and ultimately to stable form with
increasing pyrolysis temperature (600 to 800  C). During pyrolysis, above
200  C organic P (ortho-monoester) fractions changed into inorganic
forms (Ortho and Pyro-phosphate). Sodium pyro-phosphate (Na4P2O7) and
monetite (CaHPO4) were major species present in biochars, whose propor-
tions increased with pyrolysis temperature. The other stable minor forms
of P in biochars as detected by Solid state 31P NMR are Crandallite
14 A. A. KARIM ET AL.

Figure 4. P species transformation due to pyrolysis and its dominant forms present in the bio-
char fertilizers.

(CaAl3(OH)5(PO4)2), Wavellite (Al3(OH)3(PO4)25H2O), Hydroxyapatite


[Ca5(PO4)3OH], Sodium pyro-phosphates [Na2H2P2O7], Tri-sodium di-
phosphate [Na3HP2O7], Senegalite [Al2(OH)3(PO4)H2O], Variscite
(AlPO42H2O), Di-calcium phosphate dihydrate (CaHPO42H2O).
The variations in P speciation in biochars produced from solid manure
digestate of biogas plant 300–1050  C pyrolysis temperature and its effect
on P availability in soils were investigated (Bruun et al., 2017). P was dom-
inantly present as calcium phosphates in the solid digestate. The elevated
pyrolysis temperature transformed the less thermally stable forms such as
struvite and magnesium phosphates into high thermally stable apatite with
increased crystallinity even at above 1050  C. A study on biochar derived
from bone meal also revealed that with increase in pyrolysis temperature
the organic P transformed into crystalline calcium phosphates (Zwetsloot
et al., 2015). Research work undertaken by Uchimiya and Hiradate (2014)
on biochar produced from plant and manure also concluded that increase
in pyrolysis temperature gradually converts the organic phytate phosphate
into pyro-phosphate and finally to orthophosphate at further elevated tem-
peratures. On basis of above studies, scheme of P species transformation
due to pyrolysis and its dominant forms present in the biochar fertilizers is
presented in Figure 4.

2.2.3. K transformation
The transformation of K compounds into biochars initiated above 250  C
pyrolysis temperature. The water-soluble fraction of K usually transformed
into organic compounds up to 500  C; whereas between 500 to 700  C,
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY 15

silicate and volatile compounds such as K2SO4, KOH, and KCl dominated
in biochars. Furthermore, increase in temperature above 700  C leads to
loss of K from biochars through decomposition and volatilization of K-
minerals (Tian et al., 2017). Non-woody plant wastes mainly constitute of
K containing minerals specifically kalicinite and sylvite, present both on the
surface and inside voids of biochar particles. Prakongkep et al. (2015)
reported the minerals form of nutrients in biochars, produced through
slow pyrolysis from different tropical plant wastes i.e. corncob, lemon peel,
soybean cake, bamboo wood, eucalyptus wood, coconut shell, coconut fiber,
rice husk, sugar palm fiber, oil palm fruit, coffee waste, tamarind wood,
and durian shell. Most of these biochars predominantly contain calcite
(CaCO3) and kalicinite (KHCO3) minerals. Other nutrients containing min-
erals were also present in biochars such as sylvite (KCl), archerite
(KH2PO4), struvite (KMgPO4. 6H2O), chlorocalcite (KCaCl3) and pyroco-
proite (K2MgP2O7). Kalicinite is highly water-soluble mineral and thus con-
tain P in plant available form. Karim et al. (2017) also reported the
formation of kalicinite mineral in banana peduncle biochars. Formation of
Kalicinite in biochars might be through following mechanism. K mainly
exists in biomass as a dissolved salt in ionic form. Biomass pyrolysis above
400  C mainly forms KOH and KO2 due to H2O presence in volatiles/bio-
oil (Cao et al., 2016). Subsequently, KO2 may react with C or CO to form
K2CO3, which reacts with H2O molecules and CO2 to form KHCO3 (kali-
cinite mineral). Possible reactions that lead to formation of KHCO3 during
biomass pyrolysis are as follows (Equations (3)–(7)).

2K þ 2H2 O ! 2KOH þ H2 (3)

2KOH þ 2K ! 2K2 O þ H2 ðabove 400 CÞ (4)

4KO2 þ 3C ! 2K2 CO3 þ CO2 (5)

2KO2 þ CO ! K2 CO3 þ O2 (6)

K2 CO3 þ H2 O þ CO2 ! 2KHCO3 (7)

Presence of above-mentioned minerals and their dominance in biochars


strictly depends on the composition of biomass feedstocks (Figure 5).
These K-minerals are highly water-soluble and thus highlight the potential
of biochars as K-fertilizer substitute for synthetic chemical fertilizers.
Among biochars, durian shell, fruit peel, and fiber waste biochars were
relatively richer in K minerals, which indicate its more applicability as K-
fertilizers. Overall, application of these K rich biochar fertilizers could be
appropriate for sustainable organic agriculture.
16 A. A. KARIM ET AL.

Figure 5. K species transformation due to pyrolysis and its dominant forms present in the bio-
char fertilizers.

2.3. Nutrient binding mechanisms and its release


When applied to the soil, the total amount of nutrients present in biochars
necessarily does not indicate the actual amount that is released (Camps-
Arbestain et al., 2015; Singh et al., 2017). Compared to the original feed-
stock, nutrients (particularly N) in biochar are usually less bio-available.
El-Naggar et al. (2015) reported that out of 15.6% N in the original feed-
stock, only 4.5% N was retained in biochar. Higher total C/N ratio in bio-
char material leads to N immobilization and reduced bioavailability in soils
(Arif et al., 2017; Riaz et al., 2017), which could be the reason that biochar
has not contributed to improve the N crop budget (Hangs et al., 2016;
Nguyen et al., 2017).
For agricultural systems, biochars could also be a valuable method specific-
ally for P recycling. As total P amounts of animal manure produced annually
surpass world P fertilizer output, the thermochemical conversion of manure
into biochar could be a viable alternative to mineral P fertilizers (Steinfeld,
2006). P recycling from organic residues/wastes has environmental advantages
(e.g., water bodies protection) and can also provide a continuous P source for
soils (Dai et al., 2016; Manolikaki et al., 2016; Wang et al., 2015). However, the
water soluble fractions of P in biochars are reduced compared to raw biomass
due to the bonding of P to multivalent metal cations during pyrolysis (Dai
et al., 2016; Wu et al., 2012). Generally, P amount in wood-derived biochars
are very less and not bioavailable in soils, thus not suitable as a P fertilizer.
Whereas, P-rich biomass (e.g., animal manures) derived biochars contains
higher P and showed increase in its bio-availability in soils. Therefore, these
biochars could be suitable as P fertilizer (Glaser & Lehr, 2019).
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY 17

K mainly exists in plant available forms in almost all biochars, which was
due to its presence as water-soluble minerals (Ippolito et al., 2015). Although
above said biochars possessed high total N/P/K contents, the availability of
nutrients for plants depends on its chemical speciation, and need further
detailed research. Nutrients released from the biochar into the solution
depends strongly on the feedstock (Nelson et al., 2011). Furthermore, the
release of nutrient varies based on the sorption capacity of the element.
Leaching trials to study the release of nutrients from hardwood biochar car-
ried out by Angst and Sohi, (2013) reported a progressive K supply over the
growing season compared to rapid release of K. Therefore, while managing
plant nutrient supply through biochar application, the difference in the release
patterns of each of the nutrient and plant type needs proper consideration.
Application of above discussed macro-nutrients rich biochars could be
more appropriate for degraded and/or macronutrients deficit soils. For
instance, calcareous soils in arid zones have common issues of low
nutrients (specifically P) availability and organic matter (Abbaszadeh-
Dahaji et al., 2020; Bityutskii et al., 2017; Kumari et al., 2018). The decline
in organic matter adversely impacts the supply of nutrients in soil (Naeem
et al., 2017). Post biochar applications, the soil total N, available P and K
has been increased to 1.2–1.4 g/kg, 1.70–2.65 mg/kg, and 1.52–2.59 mg/kg,
respectively. The highest increase in availability of NPK in soil reported
with 2% biochar application rate. The findings also showed that pyrolysis
temperature increased the P/K availability in soils (Karimi et al., 2020).

3. Potential of biochars fertilizer for soil application


Relevant research studies are explained herewith that investigated the
potential of nutrients rich biochars (pyrolytic temperature 300–550  C) for
soil nutrients replenishment and improved plant growth (Table 1).
Limitation of these studies were high non-uniformity, and execution
through lab scale greenhouse pot experiments with different biochars and
soil-plant systems. Nutrient rich biochars showed mixed results and no spe-
cific trend about the influence of pyrolytic temperature and feedstocks on
efficacy of biochars for improving nutrient availability in soils and plant
growth was observed. Animal manures and sewage sludge derived nutrient
enriched biochars application in soils were commonly studied. Application
of NPK enriched poultry litter biochars increased the total N concentration
in sub-tropical acidic soil and foliar N in macadamia orchard plant leaves.
It also improved the P availability in that soil (Bai et al., 2015). Zolfi-
Bavariani et al. (2016) also reported enhanced P availability in calcareous
soil after application of poultry manure biochar. Treatment of clay loam
soil with poultry manure biochar increased the dry weight, P and K content
of lettuce (Lactuca sativa) and corn (Zea mays L.) (Gunes et al., 2015).
18 A. A. KARIM ET AL.

Abbasi and Anwar, (2015) research showed poultry manure biochars appli-
cation significantly increased N uptake, growth and biomass yield of wheat
(Triticum aestivum L.) in loam soil. The post-harvest examination of soil
also showed improvement in its total N content and C/N ratio than con-
trol. Application of nutrient rich cow manure biochar in sandy soil
increased the pH, water use efficiency, hydraulic conductivity, bio-available
P, and total N along with significant improvement of maize grain yield
(Uzoma et al. 2011). Subedi, Taupe, Pelissetti, et al. (2016) reported treat-
ment of manure (poultry litter and swine manure) biochars in silt loam
and sandy soils showed significant increase in yield, shoot matter dry
weight along with enhanced uptake of NPK nutrients in ryegrass plant.
However, in another study application of swine solid biochars in Mollisol
and Entisol soils showed negative impact i.e. increased in the loss of dis-
solved P and K (Novak et al., 2014).
Wastewater sludge biochar treatment of chromo-sol soil significantly
enhanced the N and P availability and increased the yield of cherry toma-
toes (Hossain et al., 2010). Application of sewage sludge biochar in tropical
clay soil significantly improved nutrient availability, uptake and growth of
maize plants (Gwenzi et al., 2016). Sewage sludge biochars also increased N
and P availability in oxisol soil and improved the growth of radish
(Raphanus sativus L.) plant (Sousa & Figueiredo, 2016). Song et al. (2014)
reported application of sewage sludge biochar in acidic soil improved
growth, dry matter, and yield of garlic (Allium sativum L.). This study
reported the heavy metal (Cu and Zn) accumulation in bulbs and roots of
garlic plants, which underscores that application of sewage/wastewater
sludge biochars for agriculture could pose risk for human health.
Application of P rich biochars produced from cattle carcasses and its appli-
cation to acidic soil significantly improved P availability leading to better
growth and dry matter of corn (Zea mays L.) plant (Ma & Matsunaka,
2013). Incorporation of K enriched biochars (prepared by treating biochars
with anaerobically digestate slurry) with slow release fertilizer slowed down
the leachability of K, but showed lower lettuce yield than commercial slow
release fertilizer (Oh et al., 2014). The decrease yield was probably due to
elevated pH of soil after biochar treatment. Research studies extrapolated
the application rate for biochars to be very high (up to 270 t/ha), which
seems to be less practically feasible keeping in view the regular requirement
of at least twice amount of dried feedstock, high production cost and other
negative effects like possibility for inducing long term alkalinity and limit-
ing nutrient availability in soils. Therefore, future nutrient enriched biochar
research is essential to focus more on high availability feedstock (e.g., for-
estry waste, crop residues), identify practically feasible application methods,
application in appropriate soil type (e.g., degraded, mine contaminated, and
forest soil) and plants (e.g., forestry and horticultural).
Table 1. Effect of nutrients enriched biochars fertilizer on soil fertility and plant growth.
Pyrolysis Quantitative detail on change “from
Biochars temperature Soil type Plant Type of study Biochar dose control to biochar treatment” References
riPoultry manure biochar 300  C Clay loam Lettuce (Lactuca Greenhouse pot 10 g/kg soil (25 t/ha) Lettuce: 1.63 g/kg P to 5.32 g/kg; 43.57 g/ Gunes
sativa) and Corn experiment kg K to 75.09 g/kg; increased lettuce et al. (2015)
(Zea mays L.) plant dry weight from 31.6 to 445.5 g
Corn: P 0.97 to 2.93 g/kg; K 52.65 to
66.56 g/kg; corn dry weight from 1023
to 1374 g
Poultry manure biochar 300  C Calcareous soil Pot 150 days 2% Soil: enhanced the availability of P from Zolfi-Bavariani
incubation 3.54 to 125 mg/kg in soil et al. (2016)
experiment
Biochars (water- 300  C Decomposed Lettuce Greenhouse pot 67.5 mg/ha Soil: increase available P (P2O5) 20.97 to Oh et al. (2014)
treatment sludge) granite soil (Lactuca sativa) experiment 71.75 mg/kg; Increase potassium from
(WS) treated with (sand at  91.8%) 0.02 to 0.15 cmol/kg; increase plant
anaerobically yield from l 2.28 to 34.22 t/ha
digestate slurry
Sewage sludge biochar 300  C Oxisol Radish (Raphanus Green house pot 50 g/kg Soil: increase P from 0.39 to 100.31 mg/kg; Sousa and
sativus L.) experiment 35 days increase K from 18 to 34.25 mg/kg; Figueiredo
increase Nitrate from 7.32 to 41.91 mg/ (2016)
kg; increase plant dry weight from 0.2
to 3 g/pot
Sewage sludge biochar 300  C Tropical Maize (Zea mays L.) 49 days pot 15 t/ha Increased total dry biomass weight from Gwenzi
clay soil experiments 20 to 80 g/pot et al. (2016)

Poultry litter biochar 400 C Silt loam and Ryegrass (Lolium 150 days Pot (1.5 kg) Seeds were sown (0.2 g/ Silt loam soil: increased plant shoot dry Subedi, Taupe,
sandy soils multiflorum L.) experiments in pot given the seed weight from around 13 to 18 g/pot Ikoyi, et al.
controlled chamber. rate of 10 g m2 ) Sandy soil: increase plant shoot dry (2016)
weight from 6 to 9 g/pot
Swine manure biochar 400  C Silt loam and Ryegrass (Lolium 150 days pot (1.5 kg) Seeds were sown (0.2 g/ Silt loam soil: increased plant shoot dry Subedi, Taupe,
sandy soils multiflorum L.) experiments in pot given the seed weight from around 13 to 16 g/pot Pelissetti,
controlled chamber rate of 10 g m2 Sandy soil: increased plant shoot dry et al. (2016)
weight from 6 to 8 g/pot
Swine solid biochars 400  C Mollisol 25-day laboratory 3.84 g/kg (ca.10 t/ha) Entisol: cumulative DP % loss increase Novak et al. (2014)
and incubation pot experiments from 1.44 to 6.03; cumulative DK% loss
entisol from increase from 6.10 to 34.19
Mollisol: cumulative DP % loss increase
from 0.30 to 0.61; cumulative DK %
loss decrease from 0.88 to 0.74
Cattle Carcasses 450  C Acidic soil Maize (Zea mays L.) 40 days pot plant 1.6 g biochar/2.4 kg Biochar improved dry matter of plant Ma and Matsunaka
biochar growth experiments in soil pot 25.6 g/pot compared to 22.0 g/pot for (2013)
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY

glasshouse superphosphate; increased P plant


uptake to 38.6 g/pot than 36.9 g/pot
for superphospate
450  C 270 t/ha Song et al. (2014)
19

(continued)
20

Table 1. Continued.
Pyrolysis Quantitative detail on change “from
Biochars temperature Soil type Plant Type of study Biochar dose control to biochar treatment” References
Sewage sludge Local Slightly Garlic (Allium 40 days pot Compared to control, biochar increase
biochars acidic (6.3 pH) sativum L.) experiments 30% dry matter yield of garlic
soil of Sanghai,
China
Poultry manure 500  C Loam soil Wheat (Triticum Greenhouse pot 30 t/ha Increase in dry matter yield of wheat from Abbasi and
biochar aestivum L.) experiment upto 11.7 to 27.0 g/pot, grain yield increase Anwar (2015)
plant maturity from 7.9 to 18.9 g/pot
A. A. KARIM ET AL.

Poultry litter 500  C Subtropical Macadamia Finding of 5-year 40 kg per tree; 10 t/ha Increased soil TN from 0.45% to 0.55%, Bai et al. (2015)
biochar acidic soil (Macadamia integrifolia L.) field study dry weight slightly high foliar N (1.34%) in biochar
treated than control 1.29%
Cow manure 500  C Sandy soil Maize (Zea mays L.) 85 days greenhouse 15 t/ha Increased maize grain yield by 150 % as Uzoma et al. (2011)
biochar pot experiments compared with the control; increased N
from 0.45 to 1.50 g/kg in soil
Sewage sludge 500  C Tropical clay soil Maize (Zea mays L.) 49 days pot 7.5 t/ha Maize growth increased by 42 % and total Gwenzi et al. (2016
biochar experiments biomass by 270% compared to control

Wastewater 550 C Chromosol Cherry tomato Glasshouse pot 10 t/ha Biochar treatment produced 64% greater Hossain et al. (2010)
sludge biochar (Lycopersicon experiments (16 wks) yield compared to control treatment
esculentum)
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY 21

3.1. Contradictory findings related to biochar application in soil


Several biochars studies have reported conflicting findings that illustrate
positive, negative or neutral impacts on the soil quality (Beiyuan et al.,
2017; Yang et al., 2019). Biochar derived from various feedstocks showed
different potential toward improving fertility of sandy and sandy loam soils
(El-Naggar et al., 2018), where rice straw biochar significantly increased the
content of N, available P, exchangeable cations, and CO2 efflux compared
to the wood and grass biochar. The results strongly depend on the type of
biochar produced from five different feedstocks (Alburquerque et al., 2014).
Authors also stated that soils treated with wheat straw and pine wood chip
biochar had a greater field potential than soils treated with other kinds of
biochar. The contrasting findings of such studies can be partly due to fac-
tors like the soil type and experimental arrangement. However, the differ-
ence in composition of each type of biochar is one of the key reasons of
contrasting results (El-Naggar et al., 2019; Rajapaksha et al., 2019). Any
biochar generated using a specified production process (e.g., pyrolysis, gas-
ification, hydrothermal carbonization), at a specified temperature and with
or without any modifications will produce a unique biochar (Igalavithana
et al., 2017; Yoo et al. 2018). Without specifying production conditions and
biochar composition it would be very problematic to generalize the func-
tion of biochar on different usages. Some papers have shown variations in
properties of biochar and soil functions on the basis of type of feedstock
and the production condition (Agegnehu et al., 2017; Ding et al., 2016;
Igalavithana et al., 2017). The application of biochar can increase the avail-
ability of soil water (Ma et al., 2016), water holding capacity (Mohamed
et al., 2016), soil aeration (Cayuela et al., 2013), soil organic C content (El-
Naggar et al., 2019), soil microbial biomass and activity (Igalavithana et al.,
2017), enzymatic activity (Awad et al., 2018), and nutrient retention and
availability, which results in fewer fertilizer needs and reduces the leaching
of nutrients (El-Naggar et al., 2019). However, some studies reported
reduce nutrient availability (Karer et al., 2013 Hussain et al., 2017) and
decrease in crop productivity following biochar application (Schmidt et al.,
2015), which could be related to a decrease in plant nutrient uptake or
mineralization of soil C (Ippolito et al. 2012). Biochar produced at high
pyrolytic temperatures (around 600  C), for example, may adsorb plant
nutrients, thereby limiting plant uptake. Moreover, the negative priming
effect (PE) induced by biochar nutrient adsorption may also lead to reduc-
tion in the availability of nutrients for plant uptake in low organic carbon
soils (Kuppusamy et al., 2016).
Biochar application for amelioration of saline/sodic soils has also been
studied, which reported contrasting results. Some studies confirmed the bio-
char application reduced the exchangeable sodium percentage (ESP) and
22 A. A. KARIM ET AL.

Sodium Adsorption ratio (SAR) in saline and sodic soils (Amini et al., 2016;
Luo et al., 2017; Sun et al., 2016; Wu et al., 2014). The different mechanisms
responsible for that could be biochar application releases exchangeable Ca to
replace Na in soil solution, organic C in biochar increases the surface charge
density leading to comparatively more Ca adsorption in soil colloids than Na,
enhancement in soil porosity due to biochar may increase the Na leaching
and decrease its concentration from soil solution (Dahlawi et al., 2018).
Generally, salt affected soils are low in NPK, which could be due to reasons:
(1) enhanced organic matter solubility increases the mineralized nutrients
leaching and loss, and (2) salinity and sodicity negatively impacts the benefi-
cial microbial population, which consequently reduced the nutrient trans-
formation and limits phyto-availability (Dahlawi et al., 2018). Biochar
application in saline and sodic soils had showed the improvement in macro-
nutrients contents (Akhtar et al., 2015; Kim et al., 2016; Lashari et al., 2013,
2015; Lin et al., 2015). This could be done by the biochar through directly
supplying the macronutrients, neutralizing pH, improving organic C content,
and indirectly supporting the growth of beneficial microbial population like
N2 fixing bacteria and phosphate solubilizing bacteria (Lashari et al, 2013).
Several research studies had also reported that biochar application could nega-
tively impact the soil by increasing ESP and SAR, as some biochars (depend-
ing on feedstocks like poultry litter) also contains considerable amount of Na
(Fernandes et al., 2019). Furthermore, several studies also reported that bio-
char application limits the nutrient bio-availability in saline and sodic soils
due to biochar induced alkalinity and nutrients (P) precipitation (Xu, Zhang,
Sun, & Shao, 2016). In this context, either biochar with less Na and high Ca/
K or biochar mixed with Ca/K rich mineral could be more suitable for appli-
cation in salt affected soils (Dahlawi et al., 2018; Lin et al., 2015). Overall,
almost all biochar research works underscores that the positive and negative
impacts of biochars application are mainly related to biomass feedstock prop-
erties, pyrolysis condition, biochar dose, and soil characteristics.

3.2. Socio-economic aspects of biochar fertilizers


In recent decades, technological advancement has provided many useful
soil productivity and ecosystem resource management techniques essential
for sustainable agriculture. But, these have rarely been discussed with pro-
ducers and perhaps other agricultural specialists, farmers and technicians
who are actually in need of some advancements (Scherr & McNeely, 2008).
This appears specifically true for biochar-based fertilizer. So, knowledge of
the socio-economic aspects of biochar production, logistics and recognition
by farmers are essential, if biochar to be adopted in agricultural practice.
While biochars have many technological advantages, it is of limited use if
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY 23

its application (how and when) not properly demonstrated and people/
farmers are unaware. Local development and distribution of crop residue
derived biochars are more sustainable than large-scale processing for farm-
ers. Farmers on-farm easy and convenient ways to turn crop residues or
any other waste biomass into biochar should practice and therefore trans-
port of biomass to any other locations are not required. However, it is dif-
ficult to convert diverse biomass into biochar by specific, low cost single
pyrolysis reactor design, as the complexity of biomass characteristics (types,
values, size, shape, density, etc.) produces several different biochars.
Modern biochar processing is an acquired know-how. Biochar products
and innovations in market as patents are released at a high price, and are
not available to farmers. In many developed and poor countries, therefore,
the industrial approach to biochar development will not be a viable choice.
Farmers should be encouraged in their fields to use biochar on a trial
basis, and experimental results with controls should be compared. Plants
grow rapidly in case of chemical fertilizer while plant grow slowly with bio-
char. Also, with 10% increase in yield, the cost of biochar fertilizers reduces
to half than that of conventional fertilizers. While farmers during field
studies are satisfied with the biochar application performance, the high cost
of biochar technology makes it difficult for its practical implementation on
a large scale (Reddy, 2014).
Biochar costs are related directly to feedstock costs, collection, transpor-
tation costs, methods of processing and co-products value (Filiberto &
Gaunt, 2013). Life cycle and economic assessment performed for biochar
by Homagain et al. (2016) reported that the total biochar production cost
includes 12% for feedstock collection, 9% for transportation cost, 36% cost
for the pyrolysis production system and 14% for land application. The feed-
stock collection depends on its availability in bulk volume and proximity
with the location of pyrolysis system. Forestry wood and agricultural (e.g.,
corncob, rice straw) wastes could be considered as high availability feed-
stock, as it can be collected in bulk from point sources and transported eas-
ily to nearby located pyrolysis system. Whereas, animal wastes (poultry/
swine manures and bone wastes) could be low available feedstock, as for
acquiring it in sufficient quantity (e.g., to produce 50 tons of biochar for
land application) they need to collected from several non-point sources
located at very large distances (within 300 km or above). Low availability
feedstock collection costs about $134,053/year, and high feedstock availabil-
ity costs $113,945/year. For low and high feedstock availability, the trans-
portation costs were $97,962/year and $83,268/year, respectively (Homagain
et al., 2016). Cost of biochars derived from green waste and waste wood
biochars are within $150 and $260/ton (Shackley et al., 2011). Biochar cost
produced from bagasse is between $50 and $200/ton (Van Zwieten et al.,
24 A. A. KARIM ET AL.

2008). The reports by United States (US) initiative assigned $500/ton broad
cost for biochar (US Biochar Initiative, 2013). Taking into account the costs
of input and use, together with the rates of biochar soil addition, an esti-
mated biochar amendment cost was $6,317/ha (Filiberto & Gaunt, 2013).
The costs of poultry litter and wastewater sludge biochars and other waste
biochars could be considerably lower, since it will prevent/remove the dis-
posal costs and pollution problems associated with those raw wastes. In
contrast to many conventional fertilizers, biochar is also projected to offer
lasting soil benefits, and is not necessary to be added to soil every year.
Here again lies a cost benefit of using biochar fertilizers over conventional/
synthetic fertilizers (Filiberto & Gaunt, 2013). This provides an important
context in which biochar applications are economically viable.

4. Environmental concerns and challenges of biochar application in soil


Some of the environmental impacts and challenges related to biochar appli-
cation in soils are discussed below:

1. C sequestration through conversion of biomass to biochar is suggested


as one of the ways of reducing global warming and climate change
(IPCC, 2014; Wang, Zheng, et al., 2013). Woolf et al. (2010) reported
biochar and its soil storage would help reduce current anthropogenic
CO2 emissions by up to 12%. By reducing usage of N fertilizer, biochar
could also minimize indirect greenhouse gas (GHG) emissions (Brassard
et al., 2016). However, all biochars are not the same and equally effi-
cient. For C sequestration, biochar should be prepared from specific
feedstock and through appropriate pyrolysis conditions (e.g., high tem-
perature) to contain higher recalcitrant and stable C. In addition, bio-
char should be produced in reactors generating few or no GHG.
2. Biochar application has been studied to treat and remediate infertile
and contaminated soils (Mandal, Sarkar, Bolan, et al., 2017; Mandal,
Sarkar, Igalavithana, et al., 2017; Xu, Lin, et al., 2018; Xu, Seshadri,
et al., 2018; Yang et al., 2018, 2019). It has been found beneficial in
immobilizing soil pollutants (Stefaniuk et al., 2017), however, several
studies also reported that some biochar products increased the availabil-
ity of harmful organic compounds, which represents a potential threat
to human health. Therefore, the content of organic toxicants in the bio-
char and their eco-toxicological impact on soil flora and fauna needs to
be evaluated (Kookana et al., 2011).
3. Biochar contains toxic Polycyclic Aromatic Hydrocarbons (PAHs),
whose concentrations and type varied based on production conditions
like residence time, temperature, slow or rapid pyrolysis (Garcia-Perez
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY 25

et al., 2008; Jose et al., 2016; Wang et al., 2017). Thus, biochar organiza-
tions such as the International Biochar Initiative (IBI) and European
Biochar Foundation have set a threshold PAHs value of 6–20 mg/kg and
12 mg/kg dry matter for basic grade and premium grade biochar,
respectively (EBC, 2013; IBI, 2012). Special care and further studies are
necessary for determining the suitable process conditions to produce
nutrient rich biochar with permissible PAHs value for its application to
agricultural soils.
4. Application of raw sewage sludge in soil is highly restricted due to the
presence of pathogens and toxic (heavy metals) contaminants. Due to
its high nutrient content and abundant availability, sewage/wastewater
sludge has been one of the most frequently used feedstocks for the pro-
duction of nutrient rich biochar fertilizers. Nearly all research studies in
literature (Khanmohammadi et al., 2015; Yuan et al., 2015) have empha-
sized that sewage sludge pyrolysis enriches the total heavy metals in bio-
chars. However, regarding the bio-available/leachable fractions of heavy
metals in sewage sludge biochars, both positive and negative results
were reported. Yuan et al. (2015) reported a decrease in the leachable
fraction of As, Cd, Cr, Cu, Pb, Ni, Zn and the bio-available fraction of
Cu, Fe, Mn, Zn in biochars of sewage sludge compared to raw sewage
sludge. Khanmohammadi et al. (2015) also reported a reduction in the
bio-available sewage sludge biochar fraction of Cu, Fe, Pb, Mn, Ni and
Zn. Khan et al. (2013) reported a decrease in As, Co, Cr, Ni, and Pb
bio-availability, but not Cd, Cu, and Zn. They also reported reduction
of As, Cr, Co, Cu, Ni, and Pb bioaccumulation but not of Cd and Zn.
Previous studies also document the presence of organic contaminants,
such as PAHs, in sludge-derived biochars rich in nutrients (Khan et al.,
2015; Waqas et al., 2015). Wang et al. (2019) highlighted about the
human health risks related to biochar associated PAH and suggested to
regulate the dose of application below 20 t/ha to manage the problem.
The toxic content in biochar and their transfer/accumulation to biota
can be restricted depending upon soil and biochar characteristics
(Malev et al., 2016). Toxic contaminants could be regulated within the
permissible limit by moderating methods and processes of biochar pro-
duction (Hale et al., 2012). Further extensive studies are needed to
know about biochar toxicants bioavailability, bio-magnification and risk
to the environment.
5. The retention of toxic contaminants in biochar fertilizers is dependent
on their chemical forms, which could be restricted by selecting proper
production conditions. Catalysts/additives (e.g., KCl) could be used with
biomass during pyrolysis to volatilize the heavy metals, or to use silica
to form silicate metals to significantly reduce their leaching capacity and
26 A. A. KARIM ET AL.

bio-availability in biochars. The effect of these additives on biochar


characteristics must be investigated, as it can also adversely affect the
availability of nutrients by influencing the surface area, porosity or fix-
ing K as silicates. Detailed, in-depth mechanistic study of heavy metals
retention, enrichment, and vaporization during sewage sludge pyrolysis
is highly essential to minimize its toxic content in biochars.
6. Biochar produced by high temperature gasification could not be suitable
for agriculture application. As, gasification of biomass at high tempera-
ture will produce a completely inert biochar and its application may
cause immobilization of nutrients in soils.

5. Future perspectives
1. Biochar fertilizers rich in P/K are generally low in N and have a higher
C/N ratio, which highlighted that the N will not be bio-available for
plant uptake. For improving crop productivity, the co-use of N fertil-
izers or compost with those biochar fertilizers is suggested in the soils.
Another method that could be explored is the post-treatment of P/K
rich biochar such as using it as a sorbent to recover N species from
wastewater and waste gases (e.g., biogas), which will enhance its N con-
tent and achieving the proper C/N ratio (below 30) for its bio-availabil-
ity in soils.
2. Current research aims mainly toward enrichment of single macronutri-
ent specifically N, P and K. But, for practical application of a single
nutrient (e.g., P) enriched biochar fertilizer, additional fertilizer (e.g., N
and K) must be applied along with it. Research on production of bio-
char complex (rich in multiple macronutrients) fertilizer is highly desir-
able to minimize the additional need of chemical fertilizer/compost for
plant growth. Production of designer biochar enriched in specific
nutrients as per the requirement of a particular soil-plant system should
also be investigated in future.
3. Biochar fertilizers containing P in form of Ca-P/Mg-P/Fe-P are not
highly soluble in water and are slowly available for plant uptake. The P
solubilizing bacteria such as Pseudomonas and Rhizobium (Rodrıguez &
Fraga, 1999) could be integrated in biochars porous structure to
enhance the availability of phosphate in soils for plant use. Likewise,
different microbes such as K and S solubilizing bacteria could possibly
be loaded in biochars for increasing the bio-availability of the nutrients
as per the requirement.
4. For controlled release of nutrients in the soil, it is recommended for
encapsulating the biochar with bio-polymers like latex, starch, cellulose,
chitosan etc. (Zhou et al. 2015). It will help in improving the nutrient
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY 27

use efficiency for precision agriculture. Biochar coating with waterborne


polyacrylate material decreased its rate of bio-degradation, which indi-
cates the controlled nutrient slow release for long time. In addition, the
coating material does not showed significant harmful effect on soil
microbial community diversity, and thus potentially environmental
friendly (Zhou et al., 2015). Biochar based slow release fertilizer pre-
pared by coating with water borne co-polymers (Polyvinyl alcohol and
polyvinylpyrrolidone). This biochar demonstrated an effective release
behavior for urea (Chen et al., 2017).
5. The incorporation of nanomaterials like nano clays and metal nanopar-
ticles in biochar fertilizer may also be done to increase its surface area
and functionality. This will enhance the capacity of the biochar for
retaining nutrients back in the soil for a longer time period, thereby
allowing slower release of nutrients in soils for a much better
plant growth. So, this would help in extending the bio-availability of
nutrients for longer time and minimize the unwanted loss of nutrients
through leaching.
6. Biochars (especially sewage sludge and manure derived) also contain
considerable amount of essential plant nutrients such as Ca, Mg, Fe, Zn,
Cu, Mn (Figueredo et al., 2017; Prasad et al., 2019). However, informa-
tion about their enrichment in biochar and their bio-availability in dif-
ferent soil system are highly scarce (Bornø et al., 2019). Future studies
could be focused on long-term field trials for assessing the heavy metals
leachability and plant uptake after biochar application in different soils
and environment conditions (e.g., various seasons-summer, winter, and
rainy). This will help to predict the risk of large scale biochar applica-
tion toward environment and human health.
7. The use of biochar fertilizers has been found to reduce the bioavailabil-
ity of micronutrients (Cu, Mn, Zn) in soils, which will cause deficiencies
and limited plant growth (Karimi et al., 2020). Thus, future studies are
necessary to overcome that limitation. For example, studies for evaluat-
ing utilization of other micronutrients materials along with biochars for
agriculture.
8. As the application of biochar is a relatively new agricultural practice,
most of the research studies have been undertaken at lab scale with dif-
ferent biochars and soils. Therefore, long term field studies are highly
essential to validate lab-scale experimental findings about detailed mech-
anism and possible negative impacts of biochar fertilizers application in
soils to reach an appropriate conclusion. It is essential to study and
determine the quantity that needs to be applied in soil. Furthermore,
there is inadequate information about the sustainability of its use in
agriculture.
28 A. A. KARIM ET AL.

6. Conclusions
Nutrients rich biochar fertilizers with desired characteristics can be pro-
duced by selecting specific feedstocks and applying appropriate production
methods. Nutrient enrichment in biochars predominantly related to charac-
teristics of feedstocks. All the three classified methods i.e. direct pyrolysis,
pre-pyrolysis and post-pyrolysis showed possibility for biochars fertilizer
production. Unlike pre- and post-treatment methods, direct pyrolysis of
nutrients rich feedstocks (cattle carcasses, bio-fermentation waste) produced
N-P-K enriched biochar fertilizers without additional treatments. For N
enrichment in biochar, low temperature pyrolysis condition i.e. 300–400  C
were suitable; in contrast it was reported to be relatively higher (about
700  C) for P and K enriched biochars. Enriched biochar fertilizers
improved fertility in diverse soils (sandy, silt-loam, clay loam, acidic soil)
and aided plant growth (e.g., maize, lettuce, rye grass etc.), which depends
on factors like total and bioavailable nutrient contents, soil properties, and
plant species. Also, long term studies to evaluate the holistic effects of bio-
char application on soil health (including soil biota) vis-a-vis plant prod-
uctivity are needed. As nutrient enriched biochars produced from
contaminated feedstocks like sewage-sludge contained toxic constituents
like heavy metals etc., therefore study about its alternative utilization like
mine reclamation and revegetation etc. should be explored. In view of con-
tinuous availability of diverse wastes and specific agro-ecological require-
ments, decentralized production of nutrients enriched biochars has brighter
scope to contribute for sustainable agriculture and socio-economic
development.

Acknowledgments
All the authors express sincere thanks to all the reviewers and editor for valuable guidance,
which helped in improving the content and quality of the manuscript. Authors also
acknowledge the support of Prof. S. Basu, Director, CSIR-IMMT, Bhubaneswar. Manish
Kumar would like to highlight the financial support provided by CSIR (OLP-86 and MLP-
75) and DST, Government of India [(DST/TDT/WM/2019/51 (g))] as GAP-335 projects
related to biochars theme at CSIR-IMMT.

References
Abbasi, M. K., & Anwar, A. A. (2015). Ameliorating effects of biochar derived from poultry
manure and white clover residues on soil nutrient status and plant growth promotion-
greenhouse experiments. PLoS One, 10(6), e0131592. https://doi.org/10.1371/journal.
pone.0131592
Abbaszadeh-Dahaji, P., Masalehi, F., & Akhgar, A. (2020). Improved growth and nutrition
of Sorghum (Sorghum bicolor) plants in a low-fertility calcareous soil treated with plant
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY 29

growth–promoting rhizobacteria and Fe-EDTA. Journal of Soil Science and Plant


Nutrition, 20(1), 31–42. https://doi.org/10.1007/s42729-019-00098-9
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. https://doi.org/10.1016/j.apsoil.2017.06.008
Akhtar, S. S., Andersen, M. N., & Liu, F. (2015). Residual effects of biochar on improving
growth, physiology and yield of wheat under salt stress. Agricultural Water Management,
158, 61–68. https://doi.org/10.1016/j.agwat.2015.04.010
Alburquerque, J. A., Calero, J. M., Barr on, V., Torrent, J., del Campillo, M. C., Gallardo,
A., & Villar, R. (2014). Effects of biochars produced from different feedstocks on soil
properties and sunflower growth. Journal of Plant Nutrition and Soil Science, 177(1),
16–25. https://doi.org/10.1002/jpln.201200652
Amini, S., Ghadiri, H., Chen, C., & Marschner, P. (2016). Salt-affected soils, reclamation,
carbon dynamics, and biochar: A review. Journal of Soils and Sediments, 16(3), 939–953.
https://doi.org/10.1007/s11368-015-1293-1
Angst, T. E., & Sohi, S. P. (2013). Establishing release dynamics for plant nutrients from
biochar. GCB Bioenergy, 5(2), 221–226. https://doi.org/10.1111/gcbb.12023
Antunes, E., Schumann, J., Brodie, G., Jacob, M. V., & Schneider, P. A. (2017). Biochar
produced from biosolids using a single-mode microwave: Characterisation and its poten-
tial for phosphorus removal. Journal of Environmental Management, 196, 119–126.
https://doi.org/10.1016/j.jenvman.2017.02.080
Arif, M., Ilyas, M., Riaz, M., Ali, K., Shah, K., Haq, I. U., & Fahad, S. (2017). Biochar
improves phosphorus use efficiency of organic-inorganic fertilizers, maize-wheat prod-
uctivity and soil quality in a low fertility alkaline soil. Field Crops Research, 214, 25–37.
https://doi.org/10.1016/j.fcr.2017.08.018
Atkinson, C. J., Fitzgerald, J. D., & Hipps, N. A. (2010). Potential mechanisms for achieving
agricultural benefits from biochar application to temperate soils: A review. Plant and
Soil, 337(1–2), 1–18. https://doi.org/10.1007/s11104-010-0464-5
Awad, Y. M., Lee, S. S., Kim, K. H., Ok, Y. S., & Kuzyakov, Y. (2018). Carbon and nitrogen
mineralization and enzyme activities in soil aggregate-size classes: Effects of biochar, oys-
ter shells, and polymers. Chemosphere, 198, 40–48. https://doi.org/10.1016/j.chemosphere.
2018.01.034
Bai, S. H., Xu, C.-Y., Xu, Z., Blumfield, T. J., Zhao, H., Wallace, H., Reverchon, F., & Van
Zwieten, L. (2015). Soil and foliar nutrient and nitrogen isotope composition (d(15)N) at
5 years after poultry litter and green waste biochar amendment in a macadamia orchard.
Environmental Science and Pollution Research International, 22(5), 3803–3809. https://
doi.org/10.1007/s11356-014-3649-2
Beiyuan, J., Awad, Y. M., Beckers, F., Tsang, D. C., Ok, Y. S., & Rinklebe, J. (2017).
Mobility and phytoavailability of As and Pb in a contaminated soil using pine sawdust
biochar under systematic change of redox conditions. Chemosphere, 178, 110–118.
https://doi.org/10.1016/j.chemosphere.2017.03.022
Biederman, L. A., & Harpole, W. S. (2013). Biochar and its effects on plant productivity
and nutrient cycling: A meta-analysis. GCB Bioenergy, 5(2), 202–214. https://doi.org/10.
1111/gcbb.12037
Bityutskii, N., Yakkonen, K., Petrova, A., & Nadporozhskaya, M. (2017). Xylem sap mineral
analyses as a rapid method for estimation plant-availability of Fe, Zn and Mn in carbon-
ate soils: A case study in cucumber. Journal of Soil Science and Plant Nutrition, 17,
1–290. https://doi.org/10.4067/S0718-95162017005000022
30 A. A. KARIM ET AL.

Bornø, M. L., M€ uller-St€


over, D. S., & Liu, F. (2019). Biochar properties and soil type drive
the uptake of macro-and micronutrients in maize (Zea mays L.). Journal of Plant
Nutrition and Soil Science, 182(2), 149–158. https://doi.org/10.1002/jpln.201800228
Brassard, P., Godbout, S., & Raghavan, V. (2016). Soil biochar amendment as a climate
change mitigation tool: Key parameters and mechanisms involved. Journal of
Environmental Management, 181, 484–497. https://doi.org/10.1016/j.jenvman.2016.06.063
Bruun, S., Harmer, S. L., Bekiaris, G., Christel, W., Zuin, L., Hu, Y., Jensen, L. S., & Lombi,
E. (2017). The effect of different pyrolysis temperatures on the speciation and availability
in soil of P in biochar produced from the solid fraction of manure. Chemosphere, 169,
377–386. https://doi.org/10.1016/j.chemosphere.2016.11.058
Camps-Arbestain, M., Amonette, J. E., Singh, B., Wang, T., & Schmidt, H. P. (2015). A bio-
char classification system and associated test methods. Biochar for Environmental
Management: Science, Technology and Implementation, 165–193.
Cantrell, K. B., Hunt, P. G., Uchimiya, M., Novak, J. M., & Ro, K. S. (2012). Impact of pyr-
olysis temperature and manure source on physicochemical characteristics of biochar.
Bioresource Technology, 107, 419–428. https://doi.org/10.1016/j.biortech.2011.11.084
Cao, W., Li, J., Lue, L., & Zhang, X. (2016). Release of alkali metals during biomass thermal
conversion. Archives of Industrial Biotechnology, 1(1), 1–3. https://strathprints.strath.ac.
uk/id/eprint/59501
Cao, X., & Harris, W. (2010). Properties of dairy-manure-derived biochar pertinent to its
potential use in remediation. Bioresource Technology, 101(14), 5222–5228. https://doi.org/
10.1016/j.biortech.2010.02.052
Cayuela, M. L., Sanchez-Monedero, M. A., Roig, A., Hanley, K., Enders, A., & Lehmann, J.
(2013). Biochar and denitrification in soils: When, how much and why does biochar
reduce N2O emissions? Scientific Reports, 3, 1732. https://doi.org/10.1038/srep01732
Chan, K. Y., & Xu, Z. (2012). Biochar: Nutrient properties and their enhancement. In
Biochar for environmental management (pp. 99–116). Routledge. https://doi.org/10.4324/
9781849770552
Chang, Y. M., Tsai, W. T., & Li, M. H. (2015). Chemical characterization of char derived
from slow pyrolysis of microalgal residue. Journal of Analytical and Applied Pyrolysis,
111, 88–93. https://doi.org/10.1016/j.jaap.2014.12.004
Chen, L., Chen, X. L., Zhou, C. H., Yang, H. M., Ji, S. F., Tong, D. S., Zhong, Z. K., Yu,
W. H., & Chu, M. Q. (2017). Environmental-friendly montmorillonite-biochar compo-
sites: Facile production and tunable adsorption-release of ammonium and phosphate.
Journal of Cleaner Production, 156, 648–659. https://doi.org/10.1016/j.jclepro.2017.04.050
Conte, P., Marsala, V., De Pasquale, C., Bubici, S., Valagussa, M., Pozzi, A., & Alonzo, G.
(2013). Nature of water-biochar interface interactions. GCB Bioenergy, 5(2), 116–121.
https://doi.org/10.1111/gcbb.12009
Cui, X., Dai, X., Khan, K. Y., Li, T., Yang, X., & He, Z. (2016). Removal of phosphate from
aqueous solution using magnesium-alginate/chitosan modified biochar microspheres
derived from Thalia dealbata. Bioresource Technology, 218, 1123–1132. https://doi.org/10.
1016/j.biortech.2016.07.072
Dahlawi, S., Naeem, A., Rengel, Z., & Naidu, R. (2018). Biochar application for the remedi-
ation of salt-affected soils: Challenges and opportunities. Science of the Total
Environment, 625, 320–335.
Dai, L., Li, H., Tan, F., Zhu, N., He, M., & Hu, G. (2016). Biochar: A potential route for
recycling of phosphorus in agricultural residues. GCB Bioenergy, 8(5), 852–858. https://
doi.org/10.1111/gcbb.12365
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY 31

Dai, Y., Zheng, H., Jiang, Z., & Xing, B. (2020). Combined effects of biochar properties and
soil conditions on plant growth: A meta-analysis. The Science of the total environment,
713, 136635. https://doi.org/10.1016/j.scitotenv.2020.136635
Ding, Y., Liu, Y., Liu, S., Li, Z., Tan, X., Huang, X., Zeng, G., Zhou, L., & Zheng, B.
(2016). Biochar to improve soil fertility. Agronomy for Sustainable Development, 36(2),
36. https://doi.org/10.1007/s13593-016-0372-z
Domingues, R. R., Trugilho, P. F., Silva, C. A., de Melo, I. C. N., Melo, L. C., Magriotis,
Z. M., & Sanchez-Monedero, M. A. (2017). Properties of biochar derived from wood and
high-nutrient biomasses with the aim of agronomic and environmental benefits. PloS
One, 12(5), e0176884. https://doi.org/10.1371/journal.pone.0176884
EBC. (2013). European Biochar Certificate–Guidelines for a Sustainable Production of
Biochar. 604 European Biochar Foundation.
El-Naggar, A. H., Usman, A. R., Al-Omran, A., Ok, Y. S., Ahmad, M., & Al-Wabel, M. I.
(2015). Carbon mineralization and nutrient availability in calcareous sandy soils
amended with woody waste biochar. Chemosphere, 138, 67–73. https://doi.org/10.1016/j.
chemosphere.2015.05.052
El-Naggar, A., El-Naggar, A. H., Shaheen, S. M., Sarkar, B., Chang, S. X., Tsang, D. C. W.,
Rinklebe, J., & Ok, Y. S. (2019). Biochar composition-dependent impacts on soil nutrient
release, carbon mineralization, and potential environmental risk: A review. Journal of
Environmental Management, 241, 458–467. https://doi.org/10.1016/j.jenvman.2019.02.044
El-Naggar, A., Lee, S. S., Awad, Y. M., Yang, X., Ryu, C., Rizwan, M., Rinklebe, J., Tsang,
D. C. W., & Ok, Y. S. (2018). Influence of soil properties and feedstocks on biochar
potential for carbon mineralization and improvement of infertile soils. Geoderma, 332,
100–108. https://doi.org/10.1016/j.geoderma.2018.06.017
El-Naggar, A., Lee, S. S., Rinklebe, J., Farooq, M., Song, H., Sarmah, A. K., Zimmerman,
A. R., Ahmad, M., Shaheen, S. M., & Ok, Y. S. (2019). Biochar application to low fertility
soils: A review of current status, and future prospects. Geoderma, 337, 536–554. https://
doi.org/10.1016/j.geoderma.2018.09.034
Fang, C., Zhang, T., Li, P., Jiang, R. F., & Wang, Y. C. (2014). Application of magnesium
modified corn biochar for phosphorus removal and recovery from swine wastewater.
International Journal of Environmental Research and Public Health, 11(9), 9217–9237.
https://doi.org/10.3390/ijerph110909217
Fang, C., Zhang, T., Li, P., Jiang, R., Wu, S., Nie, H., & Wang, Y. (2015). Phosphorus
recovery from biogas fermentation liquid by Ca-Mg loaded biochar. Journal of
Environmental Sciences (China), 29, 106–114. https://doi.org/10.1016/j.jes.2014.08.019
Fernandes, J. D., Chaves, L. H., Mendes, J. S., Chaves, I. B., & Tito, G. A. (2019).
Alterations in soil salinity with the use of different biochar doses. Revista de Ci^encias
Agrarias, 42(1), 89–98. https://doi.org/10.19084/RCA18248
Figueredo, N. A. D., Costa, L. M. D., Melo, L. C. A., Siebeneichlerd, E. A., & Tronto, J.
(2017). Characterization of biochars from different sources and evaluation of release of
nutrients and contaminants. Revista Ci^encia Agron^omica, 48(3), 3–403. https://doi.org/10.
5935/1806-6690.20170046
Filiberto, D. M., & Gaunt, J. L. (2013). Practicality of biochar additions to enhance soil and
crop productivity. Agriculture, 3(4), 715–725. https://doi.org/10.3390/agriculture3040715
Gai, X., Wang, H., Liu, J., Zhai, L., Liu, S., Ren, T., & Liu, H. (2014). Effects of feedstock
and pyrolysis temperature on biochar adsorption of ammonium and nitrate. Plos One,
9(12), e113888. https://doi.org/10.1371/journal.pone.0113888
32 A. A. KARIM ET AL.

Gao, S., DeLuca, T. H., & Cleveland, C. C. (2019). Biochar additions alter phosphorus and
nitrogen availability in agricultural ecosystems: A meta-analysis. Science of the Total
Environment, 654, 463–472. https://doi.org/10.1016/j.scitotenv.2018.11.124
Garcia-Perez, M., Wang, X. S., Shen, J., Rhodes, M. J., Tian, F., Lee, W.-J., Wu, H., & Li,
C.-Z. (2008). Fast pyrolysis of oil mallee woody biomass: Effect of temperature on the
yield and quality of pyrolysis products. Industrial & Engineering Chemistry Research,
47(6), 1846–1854. https://doi.org/10.1021/ie071497p
Gimeno-Garcia, E., Andreu, V., & Boluda, R. (1996). Heavy metals incidence in the appli-
cation of inorganic fertilizers and pesticides to rice farming soils. Environmental
Pollution, 92(1), 19–25. https://doi.org/10.1016/0269-7491(95)00090-9
Glaser, B., & Lehr, V. I. (2019). Biochar effects on phosphorus availability in agricultural
soils: A meta-analysis. Scientific Reports, 9, 9338. https://doi.org/10.1038/Fs41598-019-
45693-z
Gunes, A., Inal, A., Sahin, O., Taskin, M. B., Atakol, O., & Yilmaz, N. (2015). Variations in
mineral element concentrations of poultry manure biochar obtained at different pyrolysis
temperatures, and their effects on crop growth and mineral nutrition. Soil Use and
Management, 31(4), 429–437. https://doi.org/10.1111/sum.12205
Gwenzi, W., Muzava, M., Mapanda, F., & Tauro, T. P. (2016). Comparative short-term
effects of sewage sludge and its biochar on soil properties, maize growth and uptake of
nutrients on a tropical clay soil in Zimbabwe. Journal of Integrative Agriculture, 15(6),
1395–1406. https://doi.org/10.1016/S2095-3119(15)61154-6
Hale, S. E., Lehmann, J., Rutherford, D., Zimmerman, A. R., Bachmann, R. T.,
Shitumbanuma, V., O’Toole, A., Sundqvist, K. L., Arp, H. P. H., & Cornelissen, G.
(2012). Quantifying the total and bioavailable polycyclic aromatic hydrocarbons and
dioxins in biochars. Environmental Science & Technology, 46(5), 2830–2838. https://doi.
org/10.1021/es203984k
Hagner, M., Kemppainen, R., Jauhiainen, L., Tiilikkala, K., & Set€al€a, H. (2016). The effects
of birch (Betula spp.) biochar and pyrolysis temperature on soil properties and plant
growth. Soil and Tillage Research, 163, 224–234. https://doi.org/10.1016/j.still.2016.06.006
Hangs, R. D., Ahmed, H. P., & Schoenau, J. J. (2016). Influence of willow biochar amend-
ment on soil nitrogen availability and greenhouse gas production in two fertilized tem-
perate prairie soils. BioEnergy Research, 9(1), 157–171. https://doi.org/10.1007/s12155-
015-9671-5
Homagain, K., Shahi, C., Luckai, N., & Sharma, M. (2016). Life cycle cost and economic
assessment of biochar-based bioenergy production and biochar land application in
Northwestern Ontario. Forest Ecosystems, 3(1), 21. https://doi.org/10.1186/s40663-016-
0081-8
Hossain, M. K., Strezov, V., Chan, K. Y., & Nelson, P. F. (2010). Agronomic properties of
wastewater sludge biochar and bioavailability of metals in production of cherry tomato
(Lycopersicon esculentum). Chemosphere, 78(9), 1167–1171. https://doi.org/10.1016/j.che-
mosphere.2010.01.009
Hossain, M. K., Strezov, V., Chan, K. Y., Ziolkowski, A., & Nelson, P. F. (2011). Influence
of pyrolysis temperature on production and nutrient properties of wastewater sludge bio-
char. Journal of Environmental Management, 92(1), 223–228. https://doi.org/10.1016/j.
jenvman.2010.09.008
Huang, R., & Tang, Y. (2015). Speciation dynamics of phosphorus during (hydro)thermal
treatments of sewage sludge. Environmental Science & Technology, 49(24), 14466–14474.
https://doi.org/10.1021/acs.est.5b04140
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY 33

Hussain, M., Farooq, M., Nawaz, A., Al-Sadi, A. M., Solaiman, Z. M., Alghamdi, S. S.,
Ammara, U., Ok, Y. S., & Siddique, K. H. M. (2017). Biochar for crop production:
Potential benefits and risks. Journal of Soils and Sediments, 17(3), 685–716. https://doi.
org/10.1007/s11368-016-1360-2
IBI. (2012). Standardized product definition and product testing guidelines for biochar that
is used in soil. IBI Biochar Stand.
Igalavithana, A. D., Mandal, S., Niazi, N. K., Vithanage, M., Parikh, S. J., Mukome,
F. N. D., Rizwan, M., Oleszczuk, P., Al-Wabel, M., Bolan, N., Tsang, D. C. W., Kim, K.-
H., & Ok, Y. S. (2017). Advances and future directions of biochar characterization meth-
ods and applications. Critical Reviews in Environmental Science and Technology, 47(23),
2275–2330. https://doi.org/10.1080/10643389.2017.1421844
Inyang, M. I., Gao, B., Yao, Y., Xue, Y., Zimmerman, A., Mosa, A., Pullammanappallil, P.,
Ok, Y. S., & Cao, X. (2016). A review of biochar as a low-cost adsorbent for aqueous
heavy metal removal. Critical Reviews in Environmental Science and Technology, 46(4),
406–433. https://doi.org/10.1080/10643389.2015.1096880
IPCC. (2014). Climate change 2014, mitigation of climate change. www.mitigation2014.org
Ippolito, J. A., Novak, J. M., Busscher, W. J., Ahmedna, M., Rehrah, D., & Watts, D. W.
(2012). Switchgrass biochar affects two Aridisols. Journal of Environmental Quality,
41(4), 1123–1130. https://doi.org/10.2134/jeq2011.0100
Ippolito, J. A., Spokas, K. A., Novak, J. M., Lentz, R. D., & Cantrell, K. B. (2015). Biochar
elemental composition and factors infl uencing nutrient retention. In Biochar for envir-
onmental management (pp. 171–196). Routledge. https://doi.org/10.4324/9780203762264
Jose, M., Paneque, M., Hilber, I., Blum, F., Knicker, H. E., & Bucheli, T. D. (2016).
Assessment of polycyclic aromatic hydrocarbons in biochar and biochar-amended agri-
cultural soil from Southern Spain. Journal of Soils and Sediments, 16(2), 557–565. https://
doi.org/10.1007/s11368-015-1250-z
Joseph, S., Graber, E. R., Chia, C., Munroe, P., Donne, S., Thomas, T., Nielsen, S., Marjo,
C., Rutlidge, H., Pan, G. X., Li, L., Taylor, P., Rawal, A., & Hook, J. (2013). Shifting
paradigms: Development of high-efficiency biochar fertilizers based on nano-structures
and soluble components. Carbon Management, 4(3), 323–343. https://doi.org/10.4155/
cmt.13.23
Karer, J., Wawra, A., Zehetner, F., Dunst, G., Wagner, M., Pavel, P. B., & Soja, G. (2015).
Effects of biochars and compost mixtures and inorganic additives on immobilisation of
heavy metals in contaminated soils. Water, Air, & Soil Pollution, 226(10), 342. https://
doi.org/10.1007/s11270-015-2584-2
Karer, J., Wimmer, B., Zehetner, F., Kloss, S., & Soja, G. (2013). Biochar application to
temperate soils: Effects on nutrient uptake and crop yield under field conditions.
Agricultural and Food Science, 22(4), 390–403. https://doi.org/10.23986/afsci.8155
Karim, A. A., Kumar, M., Mohapatra, S., & Singh, S. K. (2019). Nutrient rich biomass and
effluent sludge wastes co-utilization for production of biochar fertilizer through different
thermal treatments. Journal of Cleaner Production, 228, 570–579. https://doi.org/10.1016/
j.jclepro.2019.04.330
Karim, A. A., Kumar, M., Mohapatra, S., Singh, S. K., & Panda, C. R. (2019). Co-plasma
processing of banana peduncle with phosphogypsum waste for production of lesser toxic
potassium–sulfur rich biochar. Journal of Material Cycles and Waste Management, 21(1),
107–115. https://doi.org/10.1007/s10163-018-0769-7
Karim, A. A., Kumar, M., Singh, S. K., Panda, C. R., & Mishra, B. K. (2017). Potassium
enriched biochar production by thermal plasma processing of banana peduncle for soil
34 A. A. KARIM ET AL.

application. Journal of Analytical and Applied Pyrolysis, 123, 165–172. https://doi.org/10.


1016/j.jaap.2016.12.009
Karimi, A., Moezzi, A., Chorom, M., & Enayatizamir, N. (2020). Application of biochar
changed the status of nutrients and biological activity in a calcareous soil. Journal of Soil
Science and Plant Nutrition, 20(2), 450–459. https://doi.org/10.1007/s42729-019-00129-5
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. https://doi.org/10.1021/es9031419
Kelepertzis, E. (2014). Accumulation of heavy metals in agricultural soils of Mediterranean:
Insights from Argolida basin, Peloponnese, Greece. Geoderma, 221–222, 82–90. https://
doi.org/10.1016/j.geoderma.2014.01.007
Khan, S., Chao, C., Waqas, M., Arp, H. P. H., & Zhu, Y. G. (2013). Sewage sludge biochar
influence upon rice (Oryza sativa L) yield, metal bioaccumulation and greenhouse gas
emissions from acidic paddy soil. Environmental Science & Technology, 47(15),
8624–8632. https://doi.org/10.1021/es400554x
Khan, S., Waqas, M., Ding, F., Shamshad, I., Arp, H. P. H., & Li, G. (2015). The influence
of various biochars on the bioaccessibility and bioaccumulation of PAHs and potentially
toxic elements to turnips (Brassica rapa L.). Journal of Hazardous Materials, 300,
243–253. https://doi.org/10.1016/j.jhazmat.2015.06.050
Khanmohammadi, Z., Afyuni, M., & Mosaddeghi, M. R. (2015). Effect of pyrolysis tem-
perature on chemical and physical properties of sewage sludge biochar. Waste
Management & Research, 33(3), 275–283. https://doi.org/10.1177/0734242X14565210
Kim, H.-S., Kim, K.-R., Yang, J. E., Ok, Y. S., Owens, G., Nehls, T., Wessolek, G., & Kim,
K.-H. (2016). Effect of biochar on reclaimed tidal land soil properties and maize (Zea
mays L.) response. Chemosphere, 142, 153–159. https://doi.org/10.1016/j.chemosphere.
2015.06.041
Kim, J. A., Vijayaraghavan, K., Reddy, D. H. K., & Yun, Y. S. (2018). A phosphorus-
enriched biochar fertilizer from bio-fermentation waste: A potential alternative source
for phosphorus fertilizers. Journal of Cleaner Production, 196, 163–171. https://doi.org/
10.1016/j.jclepro.2018.06.004
Kizito, S., Lv, T., Wu, S., Ajmal, Z., Luo, H., & Dong, R. (2017). Treatment of anaerobic
digested effluent in biochar-packed vertical flow constructed wetland columns: Role of
media and tidal operation. The Science of the Total Environment, 592, 197–205. https://
doi.org/10.1016/j.scitotenv.2017.03.125
Kizito, S., Wu, S., Kirui, W. K., Lei, M., Lu, Q., Bah, H., & Dong, R. (2015). Evaluation of
slow pyrolyzed wood and rice husks biochar for adsorption of ammonium nitrogen from
piggery manure anaerobic digestate slurry. The Science of the Total Environment, 505,
102–112. https://doi.org/10.1016/j.scitotenv.2014.09.096
Kookana, R. S., Sarmah, A. K., Van Zwieten, L., Krull, E., & Singh, B. (2011). Biochar
application to soil: Agronomic and environmental benefits and unintended consequen-
ces. In Advances in agronomy (Vol. 112, pp. 103–143). Academic Press. https://doi.org/
10.1016/B978-0-12-385538-1.00003-2
Kumar, A., Singh, E., Khapre, A., Bordoloi, N., & Kumar, S. (2020). Sorption of volatile
organic compounds on non-activated biochar. Bioresource Technology, 297, 122469.
https://doi.org/10.1016/j.biortech.2019.122469
Kumar, A., Singh, E., Singh, L., Kumar, S., & Kumar, R. (2020). Carbon material as a sus-
tainable alternative towards boosting properties of urban soil and foster plant growth.
Science of the Total Environment, 751, 141659. https://doi.org/10.1016/j.scitotenv.2020.
141659
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY 35

Kumari, K., Prasad, J., Solanki, I. S., & Chaudhary, R. (2018). Long-term effect of crop resi-
dues incorporation on yield and soil physical properties under rice-wheat cropping sys-
tem in calcareous soil. Journal of Soil Science and Plant Nutrition, 18, 1–40. https://doi.
org/10.4067/S0718-95162018005000103
Kuppusamy, S., Thavamani, P., Megharaj, M., Venkateswarlu, K., & Naidu, R. (2016).
Agronomic and remedial benefits and risks of applying biochar to soil: Current know-
ledge and future research directions. Environment International, 87, 1–12. https://doi.org/
10.1016/j.envint.2015.10.018
Lashari, M. S., Liu, Y., Li, L., Pan, W., Fu, J., Pan, G., Zheng, J., Zheng, J., Zhang, X., &
Yu, X. (2013). Effects of amendment of biochar-manure compost in conjunction with
pyroligneous solution on soil quality and wheat yield of a salt-stressed cropland from
Central China Great Plain. Field Crops Research, 144, 113–118. https://doi.org/10.1016/j.
fcr.2012.11.015
Lashari, M. S., Ye, Y., Ji, H., Li, L., Kibue, G. W., Lu, H., Zheng, J., & Pan, G. (2015).
Biochar-manure compost in conjunction with pyroligneous solution alleviated salt stress
and improved leaf bioactivity of maize in a saline soil from central China: A 2-year field
experiment . Journal of the Science of Food and Agriculture, 95(6), 1321–1327. https://
doi.org/10.1002/jsfa.6825
Lehmann, J., & Joseph, S. (Eds.). (2015). Biochar for environmental management: Science,
technology and implementation. Routledge.
Li, W., Feng, X., Song, W., & Guo, M. (2018). Transformation of Phosphorus in Speciation
and Bioavailability During Converting Poultry Litter to Biochar. Frontiers in Sustainable
Food Systems, 2, 20. https://doi.org/10.3389/fsufs.2018.00020
Lin, W., Lin, M., Zhou, H., Wu, H., Li, Z., & Lin, W. (2019). The effects of chemical and
organic fertilizer usage on rhizosphere soil in tea orchards. PLoS One, 14(5), e0217018.
https://doi.org/10.1371/journal.pone.0217018
Lin, X. W., Xie, Z. B., Zheng, J. Y., Liu, Q., Bei, Q. C., & Zhu, J. G. (2015). Effects of bio-
char application on greenhouse gas emissions, carbon sequestration and crop growth in
coastal saline soil. European Journal of Soil Science, 66(2), 329–338. https://doi.org/10.
1111/ejss.12225
Liu, L., Tan, Z., & Ye, Z. (2018). Transformation and Transport Mechanism of
Nitrogenous Compounds in a Biochar “Preparation–Returning to the Field” Process
Studied by Employing an Isotope Tracer Method. ACS Sustainable Chemistry &
Engineering, 6(2), 1780–1791. https://doi.org/10.1021/acssuschemeng.7b03172
Liu, L., Tan, Z., Zhang, L., & Huang, Q. (2018). Influence of pyrolysis conditions on nitro-
gen speciation in a biochar ‘preparation-application’ process. Journal of the Energy
Institute, 91(6), 916–926. https://doi.org/10.1016/j.joei.2017.09.004
Liu, X., Li, Z., Zhang, Y., Feng, R., & Mahmood, I. B. (2014). Characterization of human
manure-derived biochar and energy-balance analysis of slow pyrolysis process. Waste
Management (New York, N.Y.), 34(9), 1619–1626. https://doi.org/10.1016/j.wasman.2014.
05.027
Lou, Z., Sun, Y., Bian, S., Baig, S. A., Hu, B., & Xu, X. (2017). Nutrient conservation during
spent mushroom compost application using spent mushroom substrate derived biochar.
Chemosphere, 169, 23–31. https://doi.org/10.1016/j.chemosphere.2016.11.044
Luo, X., Liu, G., Xia, Y., Chen, L., Jiang, Z., Zheng, H., & Wang, Z. (2017). Use of biochar-
compost to improve properties and productivity of the degraded coastal soil in the
Yellow River Delta. Journal of Soils and Sediments, 17(3), 780–789. https://doi.org/10.
1007/s11368-016-1361-1
36 A. A. KARIM ET AL.

Ma, N., Zhang, L., Zhang, Y., Yang, L., Yu, C., Yin, G., Doane, T. A., Wu, Z., Zhu, P., &
Ma, X. (2016). Biochar improves soil aggregate stability and water availability in a molli-
sol after three years of field application. PLoS One, 11(5), e0154091. https://doi.org/10.
1371/journal.pone.0154091
Ma, Y. L., & Matsunaka, T. (2013). Biochar derived from dairy cattle carcasses as an alter-
native source of phosphorus and amendment for soil acidity. Soil Science and Plant
Nutrition, 59(4), 628–641. https://doi.org/10.1080/00380768.2013.806205
Malev, O., Contin, M., Licen, S., Barbieri, P., & De Nobili, M. (2016). Bioaccumulation of
polycyclic aromatic hydrocarbons and survival of earthworms (Eisenia andrei) exposed
to biochar amended soils. Environmental Science and Pollution Research International,
23(4), 3491–3502. https://doi.org/10.1007/s11356-015-5568-2
Mandal, S., Sarkar, B., Bolan, N., Ok, Y. S., & Naidu, R. (2017). Enhancement of chromate
reduction in soils by surface modified biochar. Journal of Environmental Management,
186(Pt 2), 277–284. https://doi.org/10.1016/j.jenvman.2016.05.034
Mandal, S., Sarkar, B., Igalavithana, A. D., Ok, Y. S., Yang, X., Lombi, E., & Bolan, N.
(2017). Mechanistic insights of 2,4-D sorption onto biochar: Influence of feedstock mate-
rials and biochar properties . Bioresource Technology, 246, 160–167. https://doi.org/10.
1016/j.biortech.2017.07.073
Manolikaki, I. I., Mangolis, A., & Diamadopoulos, E. (2016). The impact of biochars pre-
pared from agricultural residues on phosphorus release and availability in two fertile
soils. Journal of Environmental Management, 181, 536–543. https://doi.org/10.1016/j.jenv-
man.2016.07.012
Marshall, J. A., Morton, B. J., Muhlack, R., Chittleborough, D., & Kwong, C. W. (2017).
Recovery of phosphate from calcium-containing aqueous solution resulting from bio-
char-induced calcium phosphate precipitation. Journal of Cleaner Production, 165, 27–35.
https://doi.org/10.1016/j.jclepro.2017.07.042
Modaibsb, A. S., & Al-Sewailem, M. S. (1999, January). Heavy metals content of commercial
inorganic fertilizers marketed in the Kingdom of Saudi Arabia. Proceedings of the Fourth
International Conference on Precision Agriculture, 1745–1754. American Society of
Agronomy, Crop Science Society of America, Soil Science Society of America.
Mohamed, B. A., Ellis, N., Kim, C. S., Bi, X., & Emam, A. E. R. (2016). Engineered biochar
from microwave-assisted catalytic pyrolysis of switchgrass for increasing water-holding
capacity and fertility of sandy soil. Science of the Total Environment, 566–567, 387–397.
https://doi.org/10.1016/j.scitotenv.2016.04.169
Mukherjee, A., Zimmerman, A. R., & Harris, W. (2011). Surface chemistry variations
among a series of laboratory-produced biochars. Geoderma, 163(3–4), 247–255. https://
doi.org/10.1016/j.geoderma.2011.04.021
Naeem, M. A., Khalid, M., Aon, M., Abbas, G., Tahir, M., Amjad, M., Murtaza, B., Yang,
A., & Akhtar, S. S. (2017). Effect of wheat and rice straw biochar produced at different
temperatures on maize growth and nutrient dynamics of a calcareous soil. Archives of
Agronomy and Soil Science, 63(14), 2048–2061. https://doi.org/10.1080/03650340.2017.
1325468
Nelson, N. O., Agudelo, S. C., Yuan, W., & Gan, J. (2011). Nitrogen and phosphorus avail-
ability in biochar-amended soils. Soil Science, 176(5), 218–226. https://doi.org/10.1097/SS.
0b013e3182171eac
Nguyen, T. T. N., Xu, C.-Y., Tahmasbian, I., Che, R., Xu, Z., Zhou, X., Wallace, H. M., &
Bai, S. H. (2017). Effects of biochar on soil available inorganic nitrogen: A review and
meta-analysis. Geoderma, 288, 79–96. https://doi.org/10.1016/j.geoderma.2016.11.004
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY 37

Novak, J. M., Lima, I., Xing, B., Gaskin, J. W., Steiner, C., Das, K. C., Ahmedna, M.,
Rehrah, D., Watts, D. W., Busscher, W. J., & Schomberg, H. (2009). Characterization of
designer biochar produced at different temperatures and their effects on a loamy sand.
Annals of Environmental Science, 3, 195–206.
Novak, J. M., Spokas, K. A., Cantrell, K. B., Ro, K. S., Watts, D. W., Glaz, B., Busscher,
W. J., & Hunt, P. G. (2014). Effects of biochars and hydrochars produced from lignocel-
lulosic and animal manure on fertility of a Mollisol and Entisol. Soil Use and
Management, 30(2), 175–181. https://doi.org/10.1111/sum.12113
Oh, T. K., Shinogi, Y., Lee, S. J., & Choi, B. (2014). Utilization of biochar impregnated
with anaerobically digested slurry as slow-release fertilizer. Journal of Plant Nutrition and
Soil Science, 177(1), 97–103. https://doi.org/10.1002/jpln.201200487
Ok, Y. S., Chang, S. X., Gao, B., & Chung, H. J. (2015). SMART biochar technology—a
shifting paradigm towards advanced materials and healthcare research. Environmental
Technology & Innovation, 4, 206–209. https://doi.org/10.1016/j.eti.2015.08.003
Olsen, R. (1978). Effects of intensive fertilizer use on the human environment: A summary
review. FAO Soils Bulletin, 116, 15–33. http://www.fao.org/3/aq378e/aq378e.pdf
Peiris, C., Gunatilake, S. R., Wewalwela, J. J., & Vithanage, M. (2019). Biochar for sustain-
able agriculture: Nutrient dynamics, soil enzymes, and crop growth. In Biochar from bio-
mass and waste (pp. 211–224). Elsevier. https://doi.org/https://doi.org/10.1016/B978-0-
12-811729-3.00011-X
Prakongkep, N., Gilkes, R. J., & Wiriyakitnateekul, W. (2015). Forms and solubility of plant
nutrient elements in tropical plant waste biochars. Journal of Plant Nutrition and Soil
Science, 178(5), 732–740. https://doi.org/10.1002/jpln.201500001
Prasad, M., Chrysargyris, A., McDaniel, N., Kavanagh, A., Gruda, N. S., & Tzortzakis, N.
(2019). Plant nutrient availability and pH of biochars and their fractions, with the pos-
sible use as a component in a growing media. Agronomy, 10(1), 10. https://doi.org/10.
3390/agronomy10010010
Purakayastha, T. J., Bera, T., Bhaduri, D., Sarkar, B., Mandal, S., Wade, P., Kumari, S.,
Biswas, S., Menon, M., Pathak, H., & Tsang, D. C. W. (2019). A review on biochar
modulated soil condition improvements and nutrient dynamics concerning crop yields:
Pathways to climate change mitigation and global food security. Chemosphere, 227,
345–365. https://doi.org/10.1016/j.chemosphere.2019.03.170
Qian, T. T., & Jiang, H. (2014). Migration of phosphorus in sewage sludge during different
thermal treatment processes. ACS Sustainable Chemistry & Engineering, 2(6), 1411–1419.
https://doi.org/10.1021/sc400476j
Rajapaksha, A. U., Ok, Y. S., El-Naggar, A., Kim, H., Song, F., Kang, S., & Tsang, Y. F.
(2019). Dissolved organic matter characterization of biochars produced from different
feedstock materials. Journal of Environmental Management, 233, 393–399. https://doi.org/
10.1016/j.jenvman.2018.12.069
Reddy, S. B. N. (2014). Biochar culture: Biochar for environment and development.
MetaMeta. https://metameta.nl/wp-content/uploads/2014/08/Biocharculture-Book_20_8_
2014_finalSF.pdf
Riaz, M., Roohi, M., Arif, M. S., Hussain, Q., Yasmeen, T., Shahzad, T., Shahzad, S. M.,
Muhammad, H. F., Arif, M., & Khalid, M. (2017). Corncob-derived biochar decelerates
mineralization of native and added organic matter (AOM) in organic matter depleted
alkaline soil. Geoderma, 294, 19–28. https://doi.org/10.1016/j.geoderma.2017.02.002
Roberts, D. A., Paul, N. A., Dworjanyn, S. A., Bird, M. I., & De Nys, R. (2015). Biochar
from commercially cultivated seaweed for soil amelioration. Scientific Reports, 5, 9665.
https://doi.org/10.1038/srep09665
38 A. A. KARIM ET AL.

Rodrıguez, H., & Fraga, R. (1999). Phosphate solubilizing bacteria and their role in plant
growth promotion. Biotechnology Advances, 17(4–5), 319–339. https://doi.org/10.1016/
S0734-9750(99)00014-2
Rondon, M. A., Lehmann, J., Ramirez, J., & Hurtado, M. (2007). Biological nitrogen fix-
ation by common beans (Phaseolus vulgaris L.) increases with bio-char additions.
Biology and Fertility of Soils, 43(6), 699–708. https://doi.org/10.1007/s00374-006-0152-z
Sarkhot, D. V., Ghezzehei, T. A., & Berhe, A. A. (2013). Effectiveness of biochar for sorp-
tion of ammonium and phosphate from dairy effluent. Journal of Environmental Quality,
42(5), 1545–1554. https://doi.org/10.2134/jeq2012.0482
Scherr, S. J., & McNeely, J. A. (2008). Biodiversity conservation and agricultural sustainabil-
ity: Towards a new paradigm of ‘ecoagriculture’ landscapes. Philosophical Transactions of
the Royal Society of London. Series B, Biological Sciences, 363(1491), 477–494. https://doi.
org/10.1098/rstb.2007.2165
Schmidt, H. P., Pandit, B. H., Cornelissen, G., & Kammann, C. I. (2017). Biochar-based fer-
tilization with liquid nutrient enrichment: 21 field trials covering 13 crop species in
Nepal. Land Degradation & Development, 28(8), 2324–2342. https://doi.org/10.1002/ldr.
2761
Schmidt, H. P., Pandit, B. H., Martinsen, V., Cornelissen, G., Conte, P., & Kammann, C. I.
(2015). Fourfold increase in pumpkin yield in response to low-dosage root zone applica-
tion of urine-enhanced biochar to a fertile tropical soil. Agriculture, 5(3), 723–741.
https://doi.org/10.3390/agriculture5030723
Semida, W. M., Beheiry, H. R., Setamou, M., Simpson, C. R., Abd El-Mageed, T. A., Rady,
M. M., & Nelson, S. D. (2019). Biochar implications for sustainable agriculture and
environment: A review. South African Journal of Botany, 127, 333–347. https://doi.org/
10.1016/j.sajb.2019.11.015
Shackley, S., Hammond, J., Gaunt, J., & Ibarrola, R. (2011). The feasibility and costs of bio-
char deployment in the UK. Carbon Management, 2(3), 335–356. https://doi.org/10.4155/
cmt.11.22
Shepherd, J. G., Joseph, S., Sohi, S. P., & Heal, K. V. (2017). Biochar and enhanced phos-
phate capture: Mapping mechanisms to functional properties. Chemosphere, 179, 57–74.
https://doi.org/10.1016/j.chemosphere.2017.02.123
Shepherd, J. G., Sohi, S. P., & Heal, K. V. (2016). Optimising the recovery and re-use of
phosphorus from wastewater effluent for sustainable fertiliser development. Water
Research, 94, 155–165. https://doi.org/10.1016/j.watres.2016.02.038
Singh, B., Camps-Arbestain, M., & Lehmann, J. (Eds.). (2017). Biochar: A guide to analyt-
ical methods. CSIRO Publishing.
Singh, E., Kumar, A., Khapre, A., Saikia, P., Shukla, S. K., & Kumar, S. (2020). Efficient
removal of arsenic using plastic waste char: Prevailing mechanism and sorption perform-
ance. Journal of Water Process Engineering, 33, 101095. https://doi.org/10.1016/j.jwpe.
2019.101095
Song, W., & Guo, M. (2012). Quality variations of poultry litter biochar generated at differ-
ent pyrolysis temperatures. Journal of Analytical and Applied Pyrolysis, 94, 138–145.
Song, X. D., Xue, X. Y., Chen, D. Z., He, P. J., & Dai, X. H. (2014). Application of biochar
from sewage sludge to plant cultivation: Influence of pyrolysis temperature and biochar-
to-soil ratio on yield and heavy metal accumulation. Chemosphere, 109, 213–220. https://
doi.org/10.1016/j.chemosphere.2014.01.070
Sousa, A. A. T. C., & Figueiredo, C. C. (2016). Sewage sludge biochar: Effects on soil fertil-
ity and growth of radish. Biological Agriculture & Horticulture, 32(2), 127–138. https://
doi.org/10.1080/01448765.2015.1093545
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY 39

Spokas, K. A., Cantrell, K. B., Novak, J. M., Archer, D. W., Ippolito, J. A., Collins, H. P.,
Boateng, A. A., Lima, I. M., Lamb, M. C., McAloon, A. J., Lentz, R. D., & Nichols, K. A.
(2012). Biochar: A synthesis of its agronomic impact beyond carbon sequestration.
Journal of Environmental Quality, 41(4), 973–989. https://doi.org/10.2134/jeq2011.0069
Steckenmesser, D., Vogel, C., Adam, C., & Steffens, D. (2017). Effect of various types of
thermochemical processing of sewage sludges on phosphorus speciation, solubility, and
fertilization performance. Waste Management (New York, N.Y.), 62, 194–203. https://doi.
org/10.1016/j.wasman.2017.02.019
Stefaniuk, M., Oleszczuk, P., & Roz_ yło, K. (2017). Co-application of sewage sludge with
biochar increases disappearance of polycyclic aromatic hydrocarbons from fertilized soil
in long term field experiment. Science of the Total Environment, 599–600, 854–862.
https://doi.org/10.1016/j.scitotenv.2017.05.024
Steinfeld, H., Gerber, P., Wassenaar, T. D., Castel, V., Rosales, M., Rosales, M., & de Haan,
C. (2006). Livestock’s long shadow: Environmental issues and options. FAO.
Subedi, R., Taupe, N., Ikoyi, I., Bertora, C., Zavattaro, L., Schmalenberger, A., Leahy, J. J.,
& Grignani, C. (2016). Chemically and biologically-mediated fertilizing value of manure-
derived biochar. Science of the Total Environment, 550, 924–933. https://doi.org/10.1016/
j.scitotenv.2016.01.160
Subedi, R., Taupe, N., Pelissetti, S., Petruzzelli, L., Bertora, C., Leahy, J. J., & Grignani, C.
(2016). Greenhouse gas emissions and soil properties following amendment with
manure-derived biochars: Influence of pyrolysis temperature and feedstock type. Journal
of Environmental Management, 166, 73–83. https://doi.org/10.1016/j.jenvman.2015.10.007
Sun, J., He, F., Shao, H., Zhang, Z., & Xu, G. (2016). Effects of biochar application on
Suaeda salsa growth and saline soil properties. Environmental Earth Sciences, 75(8), 630.
https://doi.org/10.1007/s12665-016-5440-9
Takaya, C. A., Fletcher, L. A., Singh, S., Anyikude, K. U., & Ross, A. B. (2016). Phosphate
and ammonium sorption capacity of biochar and hydrochar from different wastes.
Chemosphere, 145, 518–527. https://doi.org/10.1016/j.chemosphere.2015.11.052
Tian, K., Liu, W. J., Qian, T. T., Jiang, H., & Yu, H. Q. (2014). Investigation on the evolu-
tion of N-containing organic compounds during pyrolysis of sewage sludge.
Environmental Science & Technology, 48(18), 10888–10896. https://doi.org/10.1021/
es5022137
Tian, S., Tan, Z., Kasiulien_e, A., & Ai, P. (2017). Transformation mechanism of nutrient
elements in the process of biochar preparation for returning biochar to soil. Chinese
Journal of Chemical Engineering, 25(4), 477–486. https://doi.org/10.1016/j.cjche.2016.09.
009
Tsai, W. T., Liu, S. C., Chen, H. R., Chang, Y. M., & Tsai, Y. L. (2012). Textural and chem-
ical properties of swine-manure-derived biochar pertinent to its potential use as a soil
amendment. Chemosphere, 89(2), 198–203. https://doi.org/10.1016/j.chemosphere.2012.
05.085
Uchimiya, M., & Hiradate, S. (2014). Pyrolysis temperature-dependent changes in dissolved
phosphorus speciation of plant and manure biochars. Journal of Agricultural and Food
Chemistry, 62(8), 1802–1809. https://doi.org/10.1021/jf4053385
US Biochar Initiative. (2013). http://www.biochar-us.org/
Uzoma, K. C., Inoue, M., Andry, H., Fujimaki, H., Zahoor, A., & Nishihara, E. (2011).
Effect of cow manure biochar on maize productivity under sandy soil condition. Soil Use
and Management, 27(2), 205–212. https://doi.org/10.1111/j.1475-2743.2011.00340.x
40 A. A. KARIM ET AL.

Van Zwieten, L., Meszaros, I., Downie, A., Chan, Y. K., Joseph, S. (2008, March 17). Soil
health: Can the cane industry use a bit of ‘black magic’? Australian Canegrower, 10.
https://search.informit.com.au/documentSummary;dn=144660109871696;res=IELAPA
Vanlauwe, B., Bationo, A., Chianu, J., Giller, K. E., Merckx, R., Mokwunye, U., Ohiokpehai,
O., Pypers, P., Tabo, R., Shepherd, K. D., Smaling, E. M. A., Woomer, P. L., & Sanginga,
N. (2010). Integrated soil fertility management: Operational definition and consequences
for implementation and dissemination. Outlook on Agriculture, 39(1), 17–24. https://doi.
org/10.5367%2F000000010791169998 https://doi.org/10.5367/000000010791169998
Veni, D. K., Kannan, P., Edison, T. N. J. I., & Senthilkumar, A. (2017). Biochar from green
waste for phosphate removal with subsequent disposal. Waste Management (New York,
N.Y.), 68, 752–759. https://doi.org/10.1016/j.wasman.2017.06.032
Wang, C., Wang, Y., & Herath, H. M. S. K. (2017). Polycyclic aromatic hydrocarbons
(PAHs) in biochar–Their formation, occurrence and analysis: A review. Organic
Geochemistry, 114, 1–11. https://doi.org/10.1016/j.orggeochem.2017.09.001
Wang, J., Odinga, E. S., Zhang, W., Zhou, X., Yang, B., Waigi, M. G., & Gao, Y. (2019).
Polyaromatic hydrocarbons in biochars and human health risks of food crops grown in
biochar-amended soils: A synthesis study. Environment International, 130, 104899.
https://doi.org/10.1016/j.envint.2019.06.009
Wang, Y., Hu, Y., Zhao, X., Wang, S., & Xing, G. (2013). Comparisons of biochar proper-
ties from wood material and crop residues at different temperatures and residence times.
Energy & Fuels, 27(10), 5890–5899.
Wang, Y., Lin, Y., Chiu, P. C., Imhoff, P. T., & Guo, M. (2015). Phosphorus release behav-
iors of poultry litter biochar as a soil amendment. Science of the Total Environment,
512–513, 454–463. https://doi.org/10.1016/j.scitotenv.2015.01.093
Wang, Z., Zheng, H., Luo, Y., Deng, X., Herbert, S., & Xing, B. (2013). Characterization
and influence of biochars on nitrous oxide emission from agricultural soil.
Environmental Pollution (Barking, Essex : 1987), 174, 289–296. https://doi.org/10.1016/j.
envpol.2012.12.003
Waqas, M., Li, G., Khan, S., Shamshad, I., Reid, B. J., Qamar, Z., & Chao, C. (2015).
Application of sewage sludge and sewage sludge biochar to reduce polycyclic aromatic
hydrocarbons (PAH) and potentially toxic elements (PTE) accumulation in tomato.
Environmental Science and Pollution Research International, 22(16), 12114–12123.
https://doi.org/10.1007/s11356-015-4432-8
Woolf, D., Amonette, J. E., Street-Perrott, F. A., Lehmann, J., & Joseph, S. (2010).
Sustainable biochar to mitigate global climate change. Nature Communications, 1(1),
1–9. https://doi.org/10.1038/ncomms1053
Wu, W., Yang, M., Feng, Q., McGrouther, K., Wang, H., Lu, H., & Chen, Y. (2012).
Chemical characterization of rice straw-derived biochar for soil amendment. Biomass
and Bioenergy, 47, 268–276. https://doi.org/10.1016/j.biombioe.2012.09.034
Wu, Y., Xu, G., & Shao, H. B. (2014). Furfural and its biochar improve the general proper-
ties of a saline soil. Solid Earth, 5(2), 665–671. https://doi.org/10.5194/se-5-665-2014
Xiao, F., & Pignatello, J. J. (2016). Effects of post-pyrolysis air oxidation of biomass chars
on adsorption of neutral and ionizable compounds. Environmental Science & Technology,
50(12), 6276–6283. https://doi.org/10.1021/acs.est.6b00362
Xu, G., Sun, J., Shao, H., & Chang, S. X. (2014). Biochar had effects on phosphorus sorp-
tion and desorption in three soils with differing acidity. Ecological Engineering, 62,
54–60. https://doi.org/10.1016/j.ecoleng.2013.10.027
CRITICAL REVIEWS IN ENVIRONMENTAL SCIENCE AND TECHNOLOGY 41

Xu, G., Zhang, Y., Shao, H., & Sun, J. (2016). Pyrolysis temperature affects phosphorus
transformation in biochar: Chemical fractionation and 31P NMR analysis. Science of the
Total Environment, 569–570, 65–72. https://doi.org/10.1016/j.scitotenv.2016.06.081
Xu, G., Zhang, Y., Sun, J., & Shao, H. (2016). Negative interactive effects between biochar
and phosphorus fertilization on phosphorus availability and plant yield in saline sodic
soil. Science of the Total Environment, 568, 910–915. https://doi.org/10.1016/j.scitotenv.
2016.06.079
Xu, K., Lin, F., Dou, X., Zheng, M., Tan, W., & Wang, C. (2018). Recovery of ammonium
and phosphate from urine as value-added fertilizer using wood waste biochar loaded
with magnesium oxides. Journal of Cleaner Production, 187, 205–214. https://doi.org/10.
1016/j.jclepro.2018.03.206
Xu, Y., Seshadri, B., Sarkar, B., Wang, H., Rumpel, C., Sparks, D., Farrell, M., Hall, T.,
Yang, X., & Bolan, N. (2018). Biochar modulates heavy metal toxicity and improves
microbial carbon use efficiency in soil. Science of the Total Environment, 621, 148–159.
https://doi.org/10.1016/j.scitotenv.2017.11.214
Yang, X., Igalavithana, A. D., Oh, S. E., Nam, H., Zhang, M., Wang, C. H., Kwon, E. E.,
Tsang, D. C., & Ok, Y. S. (2018). Characterization of bioenergy biochar and its utiliza-
tion for metal/metalloid immobilization in contaminated soil. Science of the Total
Environment, 640–641, 704–713. https://doi.org/10.1016/j.scitotenv.2018.05.298
Yang, X., Tsibart, A., Nam, H., Hur, J., El-Naggar, A., Tack, F. M. G., Wang, C.-H., Lee,
Y. H., Tsang, D. C. W., & Ok, Y. S. (2019). Effect of gasification biochar application on
soil quality: Trace metal behavior, microbial community, and soil dissolved organic mat-
ter. Journal of Hazardous Materials, 365, 684–694. https://doi.org/10.1016/j.jhazmat.2018.
11.042
Yao, Y., Gao, B., Chen, J., Zhang, M., Inyang, M., Li, Y., Alva, A., & Yang, L. (2013).
Engineered carbon (biochar) prepared by direct pyrolysis of Mg-accumulated tomato tis-
sues: Characterization and phosphate removal potential. Bioresource Technology, 138,
8–13. https://doi.org/10.1016/j.biortech.2013.03.057
Yao, Y., Gao, B., Inyang, M., Zimmerman, A. R., Cao, X., Pullammanappallil, P., & Yang,
L. (2011a). 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.
https://doi.org/10.1016/j.jhazmat.2011.03.083
Yao, Y., Gao, B., Inyang, M., Zimmerman, A. R., Cao, X., Pullammanappallil, P., & Yang,
L. (2011b). Biochar derived from anaerobically digested sugar beet tailings:
Characterization and phosphate removal potential. Bioresource Technology, 102(10),
6273–6278. https://doi.org/10.1016/j.biortech.2011.03.006
Ye, L., Camps-Arbestain, M., Shen, Q., Lehmann, J., Singh, B., & Sabir, M. (2020). Biochar
effects on crop yields with and without fertilizer: A meta-analysis of field studies using
separate controls. Soil Use and Management, 36(1), 2–18. https://doi.org/10.1111/sum.
12546
Yoo, J.-C., Beiyuan, J., Wang, L., Tsang, D. C. W., Baek, K., Bolan, N. S., Ok, Y. S., & Li,
X.-D. (2018). A combination of ferric nitrate/EDDS-enhanced washing and sludge-
derived biochar stabilization of metal-contaminated soils. Science of the Total
Environment, 616–617, 572–582. https://doi.org/10.1016/j.scitotenv.2017.10.310
Yuan, H., Lu, T., Huang, H., Zhao, D., Kobayashi, N., & Chen, Y. (2015). Influence of pyr-
olysis temperature on physical and chemical properties of biochar made from sewage
sludge. Journal of Analytical and Applied Pyrolysis, 112, 284–289. https://doi.org/10.1016/
j.jaap.2015.01.010
42 A. A. KARIM ET AL.

Zhang, H., Voroney, R. P., Price, G. W., & White, A. J. (2017). Sulfur-enriched biochar as
a potential soil amendment and fertiliser. Soil Research, 55(1), 93–99. https://doi.org/10.
1071/SR15256
Zhang, J., L€u, F., Zhang, H., Shao, L., Chen, D., & He, P. (2015). Multiscale visualization of
the structural and characteristic changes of sewage sludge biochar oriented towards
potential agronomic and environmental implication. Scientific Reports, 5, 9406. https://
doi.org/10.1038/srep09406
Zhang, J., & Wang, Q. (2016). Sustainable mechanisms of biochar derived from brewers
spent grain and sewage sludge for ammonia–nitrogen capture. Journal of Cleaner
Production, 112, 3927–3934. https://doi.org/10.1016/j.jclepro.2015.07.096
Zhao, L., Cao, X., Zheng, W., Scott, J. W., Sharma, B. K., & Chen, X. (2016). Co-pyrolysis
of biomass with phosphate fertilizers to improve biochar carbon retention, slow nutrient
release, and stabilize heavy metals in soil. ACS Sustainable Chemistry & Engineering,
4(3), 1630–1636. https://doi.org/10.1021/acssuschemeng.5b01570
Zhou, J., Chen, H., Thring, R. W., & Arocena, J. M. (2019). Chemical pretreatment of rice
straw biochar: Effect on biochar properties and hexavalent chromium adsorption.
International Journal of Environmental Research, 13(1), 91–105. https://doi.org/10.1007/
s41742-018-0156-1
Zhou, Z., Du, C., Li, T., Shen, Y., Zeng, Y., Du, J., & Zhou, J. (2015). Biodegradation of a
biochar-modified waterborne polyacrylate membrane coating for controlled-release fertil-
izer and its effects on soil bacterial community profiles. Environmental Science and
Pollution Research International, 22(11), 8672–8682. https://doi.org/10.1007/s11356-014-
4040-z
Zolfi-Bavariani, M., Ronaghi, A., Ghasemi-Fasaei, R., & Yasrebi, J. (2016). Influence of
poultry manure–derived biochars on nutrients bioavailability and chemical properties of
a calcareous soil. Archives of Agronomy and Soil Science, 62(11), 1578–1591. https://doi.
org/10.1080/03650340.2016.1151976
Zwetsloot, M. J., Lehmann, J., Bauerle, T., Vanek, S., Hestrin, R., & Nigussie, A. (2016).
Phosphorus availability from bone char in a P-fixing soil influenced by root-mycorrhi-
zae-biochar interactions. Plant and Soil, 408(1–2), 95–105. https://doi.org/10.1007/
s11104-016-2905-2
Zwetsloot, M. J., Lehmann, J., & Solomon, D. (2015). Recycling slaughterhouse waste into
fertilizer: How do pyrolysis temperature and biomass additions affect phosphorus avail-
ability and chemistry? Journal of the Science of Food and Agriculture, 95(2), 281–288.
https://doi.org/10.1002/jsfa.6716

You might also like