Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

Chemosphere 283 (2021) 131176

Contents lists available at ScienceDirect

Chemosphere
journal homepage: www.elsevier.com/locate/chemosphere

Biochar affects the fate of phosphorus in soil and water: A critical review
Larissa Ghodszad a, Adel Reyhanitabar a, Mohammad Reza Maghsoodi a,
Behnam Asgari Lajayer b, Scott X. Chang c, d, *
a
Department of Soil Science, Faculty of Agriculture, University of Tabriz, Tabriz, Iran
b
Health and Environment Research Center, Tabriz University of Medical Sciences, Tabriz, Iran
c
State Key Laboratory of Subtropical Silviculture, Zhejiang A&F University, Lin’an, 311300, Zhejiang, China
d
Department of Renewable Resources, University of Alberta, Edmonton, T6G 2E3, Canada

A R T I C L E I N F O A B S T R A C T

Handling editor: Patryk Oleszczuk Biochar is a promising novel material for managing phosphorus (P), a nutrient often limiting for primary pro­
duction but can also be a pollutant, in the environment. Reducing P input to the environment and finding cost-
Keywords: effective approaches to remediate P contamination are major challenges in P management. There is currently no
Eutrophication review that systematically summarizes biochar effects on soil P availability and its P removal potential from
Phosphorus dynamics
water systems. In this paper, we comprehensively reviewed biochar effects on soil P availability and P removal
Pollutant
from water systems and discussed the mechanisms involved. Biochar affects soil P cycling by altering P chemical
Pyrolysis
Soil fertility forms, changing soil P sorption and desorption capacities, and influencing microbial population size, enzyme
activities, mycorrhizal associations and microbial production of metal-chelating organic acids. The porous
structure, high specific surface area, and metal oxide and surface functional groups make biochars effective
materials for removing P from eutrophic water via ligand exchange, cation bridge, and P precipitation. Because
soil and biochar properties are widely variable, the effect of biochar on the fate of P in soil and water systems is
inconsistent among different studies. Knowledge gaps in the economic practicability of large-scale biochar
application, the longevity of biochar benefits, and the potential ecological risks of biochar application should be
addressed in future research.

1. Introduction of P in crop production systems is not sustainable, as currently much of


the P fertilizer is produced from phosphate rock, which is a
Phosphorus (P) is an essential macronutrient for plants, and P defi­ non-renewable resource; the extraction of phosphate rock will eventu­
ciency is one of the main factors limiting plant growth and crop pro­ ally become non-economical as the resource is expected to be exhausted
duction, as the majority of the world’s agricultural lands are deficient in by the end of this century (Streubel et al., 2012).
P (Wang et al., 2012). Annually, 15 million tons of P fertilizer is used to Organic matter can be applied to the soil to reduce soil P fixation and
meet plant P requirement, but only 5 to 30 percent of the applied P is increase soil P availability because organic matter and its decomposition
absorbed by crops, and the bioavailability of the remnant P is very low products (e.g., organic acids) occupy the surfaces of phosphate adsor­
due to sorption, precipitation, and microbial immobilization processes bents in soils and prevent the formation of P precipitation (Du et al.,
(P fixation) that occur in the soil (Wang et al., 2012). Such behaviors of P 2013). Adding animal manure and mulch, planting cover crops, and
in the environment cause two problems: 1) the low bioavailability of the returning plant residues to soil are methods recommended to increase
applied P means that repeated application is required to meet crop P soil organic carbon (C) storage and improve soil P availability. However,
requirement in intensively managed croplands, and 2) the repeated due to their low resistance to microbial degradation, the effect of these
application of P fertilizers causes accumulation of P in the soil, types of organic matter added to the soil decreases or disappears after a
increasing the risk of P loss through surface runoff and leaching to the relatively short period of time (Moinet et al., 2018).
groundwater, resulting in the pollution of surface water, which can Over the last couple of decades, biochar has gained worldwide
cause eutrophication. A related problem is that the repeated application attention as a novel material for environmental application. According

* Corresponding author. State Key Laboratory of Subtropical Silviculture, Zhejiang A&F University, Lin’an, 311300, Zhejiang, China.
E-mail addresses: larissaghodszad@ymail.com (L. Ghodszad), areyhani@tabrizu.ac.ir (A. Reyhanitabar), mr_maqsoodi@yahoo.com (M.R. Maghsoodi), h-asgari@
tabrizu.ac.ir (B. Asgari Lajayer), sxchang@ualberta.ca (S.X. Chang).

https://doi.org/10.1016/j.chemosphere.2021.131176
Received 19 November 2020; Received in revised form 12 April 2021; Accepted 8 June 2021
Available online 13 June 2021
0045-6535/© 2021 Elsevier Ltd. All rights reserved.
L. Ghodszad et al. Chemosphere 283 (2021) 131176

Table 1
Types of biochar production processes.
Process type Product (%)

Temperature (◦ C) Residence time (second) Gas)Synthetic gas ( Liquid (Bio-oil) Solid (Biochar) Reference

Slow pyrolysis 100–1000 Long (7200–14,400) 35 30 35 Ok et al. (2015)


350–800 35 30 35 Tomczyk et al. (2020)
Intermediate pyrolysis ~500 Moderate (10–20) 25 50 25 Ok et al. (2015)
Fast pyrolysis 300–1000 Short (2) ~25–13 ~50–75 ~12–25 Ok et al. (2015)
400–600 13 75 12 Tomczyk et al. (2020)
Gasification ~900 Moderate (10–20) 85 5 10 Ok et al. (2015)
700–1500 85 5 10 Tomczyk et al. (2020)

to the IBI (International Biochar Initiative) definition, biochar is a solid liquid (bio-oil), and gaseous (such as CO, CO2, CH4, and H2) products are
material derived from the thermochemical conversion of feedstock in an obtained (Yu et al., 2019). Depending on the pyrolysis temperature,
oxygen-limited environment. Generally, biochar is provided to meet the heating rate, and residence time, four types of pyrolysis, including slow,
following four goals: waste management, climate change mitigation, intermediate, and fast pyrolysis, and gasification, are identified
energy generation, and improvement of soil properties (Tejerina, 2010). (Table 1). The type of pyrolysis process used will affect biochar yield
The physical and chemical properties of biochars are strongly controlled (the conversion ratio of feedstock to biochar), which decreases with
by factors such as the type of feedstock and pyrolysis condition (such as increasing pyrolysis temperature (Ok et al., 2015). Also, the type, size
pyrolysis temperature, heating rate, and residence time) used (Laird and lignin content of the feedstock affect biochar yield and properties
et al., 2010). Biochar is a promising soil amendment, which is highly (Ok et al., 2015). Among the types of pyrolysis, slow pyrolysis has the
resistant to microbial degradation, and an effective adsorbent for pol­ highest yield (25–35%), and it is also the main biochar production
lutants in soil environments and aqueous solutions. process in which the feedstock is generally heated at 300–650 ◦ C for a
Many studies have examined the response of P availability to biochar few minutes to several hours at a heating rate of 10–30 ◦ C min− 1 (Kambo
application in agricultural ecosystems; however, the majority of these and Dutta, 2015). Various biochar pyrolysis methods can affect biochar
experiments involved a single biochar, pyrolysis condition, application characteristics. For example, fast pyrolysis has been shown to result in
rate, or soil type, and contradictory results are often reported. Similarly, the production of biochars with higher cation exchange capacity (CEC)
the effectiveness of biochars for P removal from wastewater has been than slow pyrolysis and gasification (Lee et al., 2010). Slow pyrolysis
variable among studies (Nobaharan et al., 2021). In addition, biochar’s produces biochars with lower mineral (compared to gasification) and
efficiency in removing P from wastewater can be low because of bio­ higher recalcitrant and fixed C (compared to fast pyrolysis) components
char’s negatively charged surfaces; modification methods have been that are more suitable for C sequestration in soil (Ok et al., 2015). Wheat
proposed to enhance biochar’s affinity for anionic pollutants such as straw-derived biochar obtained by slow pyrolysis can increase net ni­
PO3−4 in water systems (Qu et al., 2020). To the best of our knowledge, trogen mineralization, whereas biochars produced through fast pyroly­
no systematic review has been conducted to synthesize the knowledge sis led to nitrogen immobilization (Bruun et al., 2012).
on how biochar influences the biochemical cycle of P in the soil and P
removal from water systems. This review synthesizes recently published
2.2. Physical properties of biochar
data in this field, discusses mechanisms involved in biochar effects on
soil P availability and P removal from water systems, and provides
Porosity and specific surface area are the two most crucial physical
recommendations on areas for future research. With this background in
properties that strongly affect the function of biochar (bH. Li et al.,
mind, this paper 1) discusses factors affecting biochar’s various physi­
2017). Biochars with high porosity and specific surface area provide
cochemical properties, 2) assesses the effects of biochar on soil P
sites for nutrient adsorption and soil microbial colonization (Chandra
biochemical processes and soil P bioavailability, 3) reviews the sorption
and Bhattacharya, 2019; Tan et al., 2018). These two properties mark­
of P by biochars and factors influencing P sorption, including biochar
edly vary with feedstock type and pyrolysis condition. The size of pores
modification methods to improve the effectiveness of biochars for P
in biochars is highly variable and includes nano- (0.09–2 nm), micro-
sorption, and 4) recommends areas for future research, based on
(<2 nm), meso- (2–50 nm), and macropores (>50 nm) (Ok et al., 2015).
knowledge gaps identified in this review, in biochar development for
Generally, biochars produced from solid waste and animal litter exhibit
application to meet agricultural and environmental management
lower surface areas as compared to those produced from wood-based
objectives.
feedstocks and crop residues, even with higher pyrolysis temperatures,
due to the considerable variation in cellulose, lignin, and moisture
2. Biochar and factors affecting biochar properties
contents in different feedstock types (Tomczyk et al., 2020). Under the
same pyrolysis condition, wood-derived biochars have a higher surface
2.1. Biochar defined
area compared to rice husk-derived biochars (Kizito et al., 2015).
Generally, high pyrolysis temperature promotes the production of bio­
Biochar is a porous, C-rich solid product from the pyrolysis of
chars that have high porosity and a strongly developed specific surface
biomass feedstocks in an oxygen-limited environment. The main pur­
area due to the dehydration process, increased micropores, decompo­
pose of the thermochemical process of pyrolysis is to break up the
sition of organic matter (lignin and cellulose), and formation of vascular
polymeric structure in the feedstock and convert polymeric compounds
bundles or channel structure (Li et al., 2013; Rafiq et al., 2016; Zhao
into those with smaller molecular sizes (Kambo and Dutta, 2015). Bio­
et al., 2017). However, at pyrolysis temperatures higher than 600 ◦ C
char can be obtained from many different kinds of organic materials,
(depending on the feedstock type), biochar’s porous structure may be
such as crop straw, rice husk, animal manure, wood waste, sewage
blocked by tar or destroyed, leading to a decreased surface area
sludge, and other agricultural wastes (Dai et al., 2020). Biochar prop­
(Chandra and Bhattacharya, 2019; Ronsse et al., 2013; Tan et al., 2018).
erties are governed by the feedstock type and pyrolysis conditions (py­
rolysis temperature, heating rate, and residence time), with the pyrolysis
temperature as one of the most important factors affecting biochar 2.3. Chemical properties of biochar
properties. During the pyrolysis process, three phases of solid (biochar),
Biochar’s chemical properties also vary greatly with feedstock type

2
L. Ghodszad et al. Chemosphere 283 (2021) 131176

and pyrolysis condition. At the same pyrolysis temperature, biochars 3. Biochar effects on soil properties
produced from non-wood feedstocks (e.g., algae, manure, and corn
stovers) generally have a higher pH than those derived from wood The impact of biochar on soil properties has been widely studied
feedstock (Enders et al., 2012; Mukome et al., 2013). Moreover, biochars (Fig. 1); however, biochar effects on soil properties depend on the soil
produced from solid waste and animal litter show a lower C and volatile and feedstock type, and the pyrolysis condition. Morphological char­
matter content, and a higher CEC compared to biochars produced from acteristics (e.g., large surface area and highly porous structure) of bio­
wood-based feedstock and crop residues, even at high pyrolysis tem­ char can change soil hydrological and microclimatic properties, which
peratures (Tomczyk et al., 2020). Generally, biochars produced at low have been linked to changes in soil microbial community and nutrient
pyrolysis temperatures show high biochar yields and have higher dis­ cycling processes (Thies et al., 2015). Biochars produced at low tem­
solved and unstable organic C contents. In contrast, biochars produced peratures increase soil CEC (Mukherjee et al., 2011), but have minimal
at high pyrolysis temperatures have high ash and C contents, high effects on soil pH in acidic soils, and could slightly reduce the pH in
alkalinity, C stability, and resistance to microbial degradation with low alkaline soils (Laghari et al., 2015). Moreover, the application of bio­
CEC, low abundance of surface functional groups, and low O/C and H/C chars produced at low temperatures can enhance the availability of ni­
ratios (S. Li et al., 2019; Wei et al., 2019; Zhao et al., 2017). Biochars trogen, P, and potassium in acidic soils with low fertility; however,
produced under high pyrolysis temperatures have more non-volatile biochar addition to alkaline soils may decrease plant-available iron,
elements such as sodium, calcium, potassium, magnesium, and P than zinc, copper, and manganese concentrations (Gunes et al., 2014). On the
those produced under low pyrolysis temperatures (Enders et al., 2012; other hand, biochars produced at high temperatures can be efficient in
Wang et al., 2015). neutralizing soil acidity and improving soil nutrient retention (Dai et al.,
Biochar can contain a significant amount of P, which often remains in 2017).
the char and not sublimated during pyrolysis up to 700 ◦ C. The amount Although biochar application can be a useful strategy in increasing
of P in biochar varies with the feedstock type. For example, biochars agricultural productivity in different soils and cropping systems in
produced from biosolids and manure have higher P contents than those various climates, the benefits of biochar application are not always
derived from wood-based biomass (Wang et al., 2014). Phosphorus positive or realized. Biochar application may also have negative effects
concentration in biochars is typically 2 to 3 times higher than that in raw on soils and agroecosystems. For example, if the feedstock contains
material due to the cleavage of organic P bonds and C sublimation heavy metals such as chromium and copper, these heavy metals will
during pyrolysis (Gul et al., 2015), as the C contained in the feedstock is remain in biochar and enter the soil system when biochar is applied.
lost faster than P during pyrolysis. The P concentration in biochar varies However, when agricultural waste material is used for biochar produc­
from 0.13 to 42.79 g kg− 1 in 38 types of biochars produced from various tion, the risk of heavy metal contamination to the soil is low. In addition,
feedstock types and under various pyrolysis conditions (Gul et al., biochar may contain many volatile organic compounds, and some of
2015). The gradual release of P from biochar contributes to the increase them, such as polycyclic aromatic hydrocarbons (PAHs), poly­
in soil available P. Phosphorus compounds in biochar encompass a chlorinated dibenzodioxins (PCDDs), and polychlorinated biphenyl
complex mixture of P species, including amorphous, semi-crystalline, (PCBs), are toxic (Hale et al., 2012; Hilber et al., 2012); however, the
and crystalline components (whitlockite, dehydrated struvite, and hy­ effect of these volatile organic compounds on plant and microbial
droxyapatite) with organic components (Enders et al., 2017). growth is not fully understood. Therefore, it is necessary to study the
bioavailability of these compounds after biochar addition to soils.

Fig. 1. Effect of biochar on soil physical, chemical, and microbiological properties (DeLuca et al., 2015; El-Naggar et al., 2019; Gao and DeLuca, 2016).

3
L. Ghodszad et al. Chemosphere 283 (2021) 131176

4. Effects of biochar application on soil P and low P availability in soils increase the need for repeated P fertilizer
application to meet the crop demand for P; rock phosphates and other
4.1. Soil P chemistry sources of P fertilizers are a non-renewable and dwindling resource. On
the other hand, repeated P fertilization and P accumulation in arable
When P fertilizers are applied to improve soil productivity, more soils can increase P runoff and leaching and increase the risk for
than 80% of the applied P becomes unavailability for plant uptake due to eutrophication in water bodies to occur (Xu et al., 2019). Hence, P
sorption, precipitation, and microbial immobilization processes, which fertilization is considered a double-edged sword in intensively managed
can remove phosphate ions from the soil solution (Zhu et al., 2018). terrestrial ecosystems (Tamburini et al., 2014).
Phosphate ions interact with soil constituents, mainly Fe and Al Biochar can alter soil P chemistry by interaction with soil organic
oxy-hydroxides, carbonates, as well as silicate clays, to decrease P and mineral components and by changing P chemical forms and the
availability in soils (Antoniadis et al., 2016). The chemical fixation of P soil’s P sorption and desorption capacity (Hong and Lu, 2018; Xu et al.,

Table 2
Biochar effects on soil P chemistry.
Feedstock Pyrolysis condition Application rate (% rates Result Reference
were based on w/w)

Mixed hardwood (primarily oak Slow pyrolysis (traditional kilns) 0, 0.5, 1 and 2% The 2% addition decreased total dissolved P leaching by 69% in Laird et al.
and hickory) soils amended with manure (2010a)
Pepperwood and peanut hull Pyrolysis at 600 ◦ C 2% Peanut hull biochar reduced phosphate in leachates by 20.6%. Yao et al. (2012)
However, pepperwood biochar increased phosphate release from
the soil
A mixture of three biochars Fast pyrolysis between 450 and 26% Biochar decreased P sorption capacity by 55% in degraded Morales et al.
(sawdust, elephant grass, and 500 ◦ C with 8 s of residence time tropical mineral soils. The soil/biochar mixture had a common P (2013)
sugar cane leaves) desorption curve but also buffered the soil solution around 0.2
mg L− 1
Corn stover, Pandrosa pine wood Fast pyrolysis at 650 ◦ C, 18 min of 4% Biochar decreased P sorption and increased P availability in acidic Chintala et al.
residue, switchgrass residence time soils, with opposite results for calcareous soils, especially with (2014)
alkaline corn stover and switchgrass biochars
1
A mixture of Norway spruce and Pyrolysis at 550–600 ◦ C, for 10–15 0, 15 and 30 t ha− Biochar had a very low P sorption affinity and did not enhance P Soinne et al.
Scots pine wood chips min sorption (2014)
Birchwood Pyrolysis at 500 ◦ C 20 t ha− 1 Biochar increased P release compared to reference soils Kumari et al.
(2014)
Wheat straw Pyrolysis at 350–550 ◦ C 0, 1, 5 and 10% P sorption increased with increasing application rate in acidic soil Xu et al. (2014)
but slightly decreased in alkaline soil. P desorption increased at
all application rates
Maize straw Pyrolysis at 400 ◦ C, 1.5-h residence 0, 2, 4 and 8% The 8% addition rate increased Olsen-P from 3 to 46 mg P kg− 1 in Zhai et al. (2015)
time acidic soil and from 13 to 137 mg P kg− 1 in alkaline soils
Mallee (Eucalyptus polybractea) Pyrolysis at 720 ◦ C, 20 min of 5% Biochar addition increased the % P retention in acidic soil by Zhang et al.
residence time ~17% (2016)
Sewage sludge 300, 400, 500, 600 and 700 ◦ C for 3 1% Sewage sludge biochars produced at 500 and 700 ◦ C reduced the Yuan et al. (2016)
h leaching of phosphate from a Typic Plinthudult soil
Hardwood and poultry litter Slow pyrolysis at 650–700 ◦ C for 1, 2 and 5% Biochar addition increased the maximum P retention capacity of Dari et al. (2016)
hardwood and at 700 ◦ C for poultry two acidic soils with increasing biochar application rate
litter, for 1 h
1
Coffee husk Pyrolysis at 350 and 500 ◦ C, Biochar produced at 500 ◦ C applied at 15 t ha− 1 increased P
0, 5, 10 and 15 t ha− Dume et al.
residence time of 3 h availability in acidic soil with low available P, but increased P (2017)
sorption and decreased P availability in a calcareous soil at higher
application rate and pyrolysis temperature
Eucalyptus globulus Pyrolysis at 450 C with a residence 0, 5, 10, 20 and 35%

P sorption was lower in acidic soils treated with biochar. 35% Martínez et al.
time of 3 h biochar application rate increased cumulative desorption by 85% (2017)
Pinewood nd wheat straw Wood: gasification at 2% Straw biochar did not affect P sorption, whereas wood biochar Bornø et al.
1000–1200 ◦ C. Straw: pyrolysis at enhanced P sorption in soils with low and intermediate P sorption (2018)
700 ◦ C capacities yet decreased P sorption in acidic soils with high P
sorption capacities
Wood chip Pyrolysis at 500 ◦ C 5% Biochar facilitated P retention in acidic soils but decreased it in Chen et al.
alkaline soils (2018a)
Cow dung impregnated with Pyrolysis at 600 C for 1 h

1% Biochar lowered P loss from leaching by 89% and enhanced Chen et al.
magnesium available P content in the soil surface layer by 3.5-fold (2018b)
Rice husk, bamboo, deciduous Slow pyrolysis at 450 ◦ C for 2 h 0, 1 and 2% Available P in biochar amended soils varies greatly with soil and Hong and Lu
tree leaves, reeds and rice biochar feedstock type. Application of rice straw biochar (2018)
straw produced at 450 ◦ C resulted in the highest available P
Rice straw Pyrolysis at 600 ◦ C for 2 h 0, 1, 3 and 5% Within a certain range of P concentration in the sorption Liu et al. (2018)
equilibrium solution, the amount of P sorbed by the three acidic
soils derived from a granite parent material markedly reduced
with increasing biochar application rate
Corn cob and rice husk Pyrolysis at 300, 450 and 650 ◦ C 1% Biochars produced at 300 and 450 ◦ C decreased P sorption and Eduah et al.
enhanced P availability, especially in acidic soils, but might (2019)
enhance P retention in neutral and alkaline soils in a short-term,
reducing P de-sorbability
Wood chips Pyrolysis at 500 C

0, 2 and 4% Biochar decreased P sorption but enhanced P binding strength in a Amarakoon et al.
sandy loam soil (2019)
Ecoera: a commercial product Slow pyrolysis at 600 ◦ C and then Applied as a 3-cm layer, Decreased P loss from medium- to high-P organic soils but was Riddle et al.
manufactured in Sweden coated with magnesium (hydr)oxide 27 cm below the soil less useful in mineral soils (2019)
surface
Maize stalk Pyrolysis at 350 and 600 ◦ C, for 2 h 1% Biochars reduced available soil P through P adsorption Li et al. (2020a,
2020b, 2020c)

4
L. Ghodszad et al. Chemosphere 283 (2021) 131176

2014). Understanding the effect of biochar application on soil P reten­ (by providing food source) and indirectly (by altering the microenvi­
tion and release can help improve P management to boost soil P ronment) increase microbial populations by creating high internal sur­
bioavailability while mitigating off-site P transport. However, results on face areas due to its pore structure and increase the capability to adsorb
biochar effects on soil P chemistry have been inconsistent and are organic matter, which provides a more diverse habitat for microbes
dependent on the type of feedstock, pyrolysis condition, application rate (Winsley, 2007). Moreover, increased soil water retention following
as well as the physical and chemical properties of the soil (Table 2). biochar application can increase microbial activities. Therefore, biochar
Little research has been done to understand the mechanisms of biochar can alter soil P bioavailability by influencing microbial population size,
effects on soil P chemistry. Generally, the mechanisms that can lead to enzymatic activity, microbial production of metal-chelating organic
increased P availability in biochar treated soils are: acids, and mycorrhizal associations (Gao et al., 2019), depending on the
soil property, biochar type, biochar application rate, and duration of
1. The biochar surface is often negatively charged and can directly experimentation (Table 3). For example, the addition of biochar can
adsorb cations such as Ca2+, Al3+, Fe2+, and Fe3+, resulting in a increase P mineralization by increasing microbial biomass by ~3 times
decrease in P sorption and precipitation (Xu et al., 2014). and phosphatase activity by ~19% 20 days after application of ~3.5 t
2. Phosphorus precipitation reactions are strongly influenced by soil pH ha− 1 of biochar produced from a water hyacinth (Eichornia crassipes)
because the activities of ions capable of precipitation with P (e.g., feedstock via slow pyrolysis at 300 ◦ C (Masto et al., 2013). Moreover,
Ca2 +, Fe2 +, Fe3+ and Al3+) are pH-dependent. Therefore, biochar the increase in soil pH following the addition of biochar, particularly
can improve P availability by modifying soil pH and affecting the those with high alkalinity, is another possible mechanism that increases
activity of these cations (DeLuca et al., 2009; Gao et al., 2019). alkaline phosphatase activities and, subsequently, P availability. For
3. Increased soil pH upon biochar addition can reduce P sorption in the instance, the addition of swine manure biochar produced at 400 ◦ C at a
soil due to stronger anionic repulsion (Chen et al., 2018a). rate of 1.5% to a clay loam and a silt loam soil increased alkaline
4. Dissolved organic matter released from biochar in the form of phosphomonoesterase activity by 28.5 and 95.1%, respectively, coin­
organic ligands can form organo-metal complexes with iron and ciding with the increase of soil pH from 6.9 to 7.5 and 5.2 to 5.8,
aluminum oxides. As a result, the effective amount of these oxides in respectively. However, acid phosphomonoesterase activities decreased
the soil is reduced, and subsequently, the adsorption of P on the by 18.6 and 34.0% in the clay loam and silt loam soils, respectively,
surfaces of these oxides is reduced (Liu et al., 2018). upon the swine manure biochar addition at the same rate (Jin et al.,
5. Biochars are a source of available P in soils. The gradual release of P 2016).
from biochar can increase soil P availability. Therefore, biochar has Phosphate solubilizing bacteria, accounting for 1–50% of the total
the potential to be used as a slow-release P fertilizer. For example, microbial population, are a key biological contributor to enhance soil P
poultry manure biochar can be a potential source of P and could availability for plants. Biochar can improve P availability in the soil by
increase extractable P in Ultisols, Oxisols, and Entisols 100 days after providing habitat and carbon supply to phosphate solubilizing bacteria,
application (Mendes et al., 2015). However, the amount and form of which can solubilize the insoluble complexed P in the soil (Siddiqui
available P are dependent on biochar type (Zhang et al., 2016). et al., 2016). The phosphate solubilizing bacteria, which excrete organic
6. The higher labile organic C concentration in biochars produced acids (i.e., citric acid, lactic acid, gluconic acid, and oxalic acid), can
under low temperature can reduce P sorption in soils due to solubilize the insoluble P. These organic acids can directly dissolve the
competitive reactions between P and low molecular weight organic phosphate mineral by anion exchange of PO3− 4 by acid anion or can
acids and anions (Eduah et al., 2019; Schneider and Haderlein, chelate Al3+ and Fe2+, and Fe3+ cations (Narula et al., 2000). For
2016). example, a biochar produced at a lower temperature (400 ◦ C) was more
7. Biochars, particularly hydrophobic or charged (negatively/posi­ vulnerable to a phosphate solubilizing bacteria (Pseudomonas putida)
tively) biochars, can adsorb organic molecules (such as complex than that produced at a higher temperature (700 ◦ C), as P species in the
proteins, phenolic acids and carbohydrates) that can act as chelates biochar produced at 400 ◦ C have a lower degree of polymerization and a
of metal ions that otherwise precipitate P. In other words, organo- poorer crystal structure than those in the biochar produced at 700 ◦ C
biochar or organo-mineral-biochar complexes are created over (Qian et al., 2019). Application of a rice husk biochar produced at 400 ◦ C
time, increasing P solubility and availability (Gao and DeLuca, increased the activities of phosphate solubilizing bacteria and led to
2016). increasing phosphorous availability (Liu et al., 2017). In addition, a
combination of sewage sludge biochar and the phosphate-solubilizing
Increased P sorption and reduction of its availability upon biochar Penicillium aculeatum can increase plant biomass and P uptake by
addition to soils, on the other hand, have been documented and have wheat and thus increase the P fertilizer value of biochar (Efthymiou
been attributed to the increased sorption sites, particularly with the et al., 2018). Holm oak-derived biochar produced at 480 ◦ C can also
addition of biochars with high specific surface areas and with high basic increase rock phosphate solubilization by Aspergillus niger by increasing
cation concentrations, especially those pyrolyzed at high temperatures; organic acid production and alleviating fluoride toxicity (Mendes et al.,
such biochars can provide a large number of positive charges and basic 2014).
cations such as Ca2+ and Mg2+ to the soil that precipitates P ions as Arbuscular mycorrhizal fungi (AMF) are one of the most important soil
calcium or magnesium phosphates (Chintala et al., 2014; Xu et al., microbial groups in terrestrial ecosystems. In exchange for sugars, AMF
2014). The contradictory effects of biochars on P behavior in different benefits their hosts by improving plant-water relations, increasing access to
soils emphasize the need to consider both soil and biochar properties immobile nutrients, especially P, and enhancing resistance to pathogens
when biochar is applied as a soil amendment, especially when it is (Warnock et al., 2010). Moreover, with plants and fungi, both secreting
intended to provide plant P requirements and to control eutrophication. phosphate solubilizing organic acids and extracellular phosphatases can
enhance plant P uptake (Gul and Whalen, 2016). Biochar can enhance AMF
4.2. Soil biological processes and P availability growth probably by providing a shelter or mineral nutrients contained in or
adsorbed on its surface (Ogawa and Okimori, 2010; Hammer et al., 2014).
Organic P needs to be mineralized to inorganic forms to become The AMF hyphae can also access microsites within the biochar that are too
available for plants. Microbial activities, including enzyme activities, small (<10 mm) for most of the plant roots to enter and can mediate plant P
are the most important determinants of P mineralization (Böhme et al., uptake from the biochar (Hammer et al., 2014). Results on the interaction
2005). Organic P hydrolysis is carried out by extracellular enzymes (i.e., between biochar and AMF are inconsistent (Table 3), which demands more
phospholipase, phosphatase, and phytase) produced by soil microor­ detailed research, particularly in realistic soil environments, to better un­
ganisms and plant roots (Gul and Whalen, 2016). Biochars can directly derstand the underlying mechanisms.

5
L. Ghodszad et al. Chemosphere 283 (2021) 131176

Table 3
Biochar affects soil P bioavailability by altering soil biological activities.
Feedstock Pyrolysis condition Application rate Result Reference

Lodgepole pine wood Pyrolysis at 600 C for 1 h



0, 0.5, 1, 2 and 4% Biochar application at 2 and 4% reduced arbuscular mycorrhizal fungi Warnock et al.
stock (w/w) (AMF) abundance in roots by 58 and 73%, respectively, but not in soils; (2010)
soil P availability was reduced by 28 and 34%, respectively
Peanut shell pellets Pyrolysis at 360, 400 and 10% (v/v) Biochar produced at 360 ◦ C enhanced P availability by 101% but reduced Warnock et al.
430 ◦ C, for 5 min AMF root colonization and extra-radical hyphal lengths by 74 and 95%, (2010)
respectively
Mango trunks and The feedstock was stacked, 0, 11.6, 23.2 and Field application of biochar at 23.2 and 116.1 t C ha− 1 enhanced P Warnock et al.
branches covered with soil and grass, and 116.1 t C ha− 1 availability by 163 and 208%, respectively, and reduced AMF abundance (2010)
ignited in soils by 43 and 77%, respectively
A mixture of spruce and 550 ◦ C for 12 h The higher nutrient content of the biochar enhances its fungal Hammer et al.
pine wood colonization. Hyphal surface contact with the produced biochar increases (2014)
P transfer to the host plant six-fold.
coniferous wood chips 500 ◦ C for 5 h 0 and 5% (v/v) P in the lettuce plant increased with AMF and biochar addition. Hammer et al.
(2015)
Birch wood 500 ◦ C 0.74% (w/w) AMF colonization of roots does not respond significantly to wood-derived Liu et al. (2017b)
biochar addition and plant P uptake decreased.
− 1
Wheat straw Pyrolysis between 350 and 0, 20 and 40 t ha Both 20 and 40 t h− 1 addition enhanced alkaline phosphatase activity but Chen et al. (2013)
550 ◦ C did not impact acid phosphatase activity
Eichornia biomass Pyrolysis at 300 ◦ C, 30 min of 0, 0.1, 0.3, 0.5, 1 Application at 2% increased alkaline phosphatase (22.8) and acid Masto et al.
residence time and 2% (w/w) phosphatase activities (32%) (2013)
Maize straw Pyrolysis at 400 ◦ C, and 1.5-h 0, 2, 4 and 8% (w/ Biochar decreased acid phosphomonoesterase activity in acidic soil and Zhai et al. (2015)
residence time w) alkaline phosphomonoesterase activity in an alkaline soil
Bamboo and rice straw Bamboo pyrolyzed at 750 ◦ C for 0, 1 and 5% (w/w) Biochars did not affect acid phosphatase activity Yang et al. (2016)
3 h and rice straw at 500 ◦ C for
30 min
Peanut shell Pyrolysis of acid-activated 0, 2.5, 5 and 10% The highest acid and alkaline phosphatase activities lasted up to 30 days Bhaduri et al.
peanut shell at 300 ◦ C for 2 h (w/w) of incubation with the lowest application rate (2.5%) (2016)
Salicornia bigelovii Pyrolysis at 350 ◦ C for 6 h 5% (w/w) Biochar increased alkaline phosphatase activities by 1.5-fold but only Al Marzooqi and
increased acid phosphatase activities on days 0 and 8 of the incubation Yousef (2017)
Rice husk Pyrolysis at 400 ◦ C for 0.5 h 0, 10, 20 and 40 t The 20 t ha− 1 treatment increased soil microbial biomass P and Liu et al. (2017a)
ha− 1 phosphatase activities the most
Wheat straw Pyrolysis at 450 ◦ C 0, 20, 40 and 60 t Alkaline phosphatase enzyme activities increased by 2–147% as a Cui et al. (2019)
ha− 1 function of increasing biochar addition rate
Moso bamboo chips Pyrolysis at 600 ◦ C 0, 20 and 40 t ha− 1 Addition at 40 t ha− 1 decreased acid phosphatase activity by 17% Peng et al. (2019)
Mentha arvensis Pyrolysis at 450 ◦ C 0, 2 and 4% (w/w) Biochar increased alkaline and acidic phosphatase activities, especially Nigam et al.
the former (2019)
Maize stalk Slow pyrolysis at 400 and 1% (w/w) Biochar enhanced alkaline phosphatase activity, with the effect greater in Khadem and
600 ◦ C, for 2 h sandy loam than in clayey soils. Increasing pyrolysis temperature Raiesi (2019)
negatively affected soil microbial activity and biomass, and subsequently,
the alkaline phosphatase activity
Rice straw Pyrolysis temperature 5% (w/w) Biochar addition decreased acid and neutral phosphatase activities Liu et al. (2019)
450–550 ◦ C, for 45 min
Rice husk Pyrolysis at 350 ◦ C 0, 3, 6 and 12 t Biochar increased alkaline phosphatase activity in the 0–0.10 m soil layer Oladele (2019)
ha− 1 after two years of field application, especially at 12 t ha− 1
Banana peel Pyrolysis at 350 C for 3 h

0, 1 and 2% (w/w) Biochar increased alkaline phosphatase activity by 12 and 6% for the Sial et al. (2019)
application rates of 1% and 2%, respectively
Rice husk Pyrolysis at ~450 ◦ C for 3 h 0, 0.5, 2 and 8% Biochar (applied at 8%) markedly increased neutral phosphatase Wang et al.
(w/w) activities by 2.3-fold compared to the control (2019)
Rice straw Pyrolysis at 380–400 ◦ C with a 0, 1, 2 and 4% (w/ Biochar applied at 1, 2 and 4% decreased alkaline phosphatase activity by Mukherjee et al.
residence time of 2.5–3.0 min w) 19, 34 and 50%, respectively, compared to the control (2019)
A mixture Eucalyptus Pyrolysis temperature between 0, 1, 2, 3, 4 and 5% Alkaline phosphatase activities were decreased with an increasing biochar Silva et al. (2020)
urophylla and 350 and 400 ◦ C (v/v) application rate
E. saligna wood
Cornstalk Pyrolysis at 250, 550 and 0.1% (w/w) Biochar produced at 550 ◦ C increased the activity of phosphate- Tao et al. (2020b)
850 ◦ C for 2 h solubilizing bacteria

5. Removal of P from water systems aquatic ecosystems, reduces biodiversity, and causes fish death, algal
growth, and economic loss. Therefore, it is essential to use appropriate
5.1. Biochar as an adsorbent technology to eliminate these elements and prevent them from entering
aquatic systems. Different physical, chemical, and biological methods
The increasing discharge of industrial, agricultural, and urban ef­ have been used to control and remove these elements from aquatic
fluents into surface waters has caused P pollution, which may result in systems. Also, previous studies have recognized adsorption as a common
eutrophication and damage to the receiving waters (Drizo et al., 1999). and widely used technique owing to its cost-efficient operation, high
The minimum phosphate threshold concentration for the induction of removal efficiency, and minimal generation of secondary byproducts
eutrophication is 0.02 mg L− 1 (Yao et al., 2013a). Today, eutrophication (Tehrani et al., 2020). In general, the use of chemical methods has its
or enrichment of surface waters is a serious environmental crisis in the disadvantages due to the high consumption of chemicals and the pos­
world (Yin et al., 2017). Due to increasing population growth and ur­ sibility of creating new pollutants in the environment.
banization, municipal, industrial and agricultural waste production has The disadvantages of biological methods can be the difficulty of
also increased. Wastewater discharge and excessive use of pesticides and implementation, high microbial sensitivity to pH in aquatic systems, and
fertilizers lead to the enrichment of water systems with nitrogen and P, potentially high cost. Physical methods, such as electrodialysis, reverse
thereby intensify the eutrophication of water bodies. This disturbs osmosis, and ion exchange for P removal, usually do not provide

6
L. Ghodszad et al. Chemosphere 283 (2021) 131176

Fig. 2. Mechanism of the formation of outer- and inner-sphere complexes of phosphate-metal on biochar surfaces (modified from Shepherd et al., 2017).

satisfactory results (Park et al., 2015). However, adsorbents such as also react with biochar surfaces via hydrogen bonds and outer-sphere
aluminum oxide, fly ash, zeolite, activated C, loess, and synthetic ma­ electrostatic interactions involving cations. This type of phosphate
terials have been used to remove inorganic pollutants from aquatic sorption by biochar is poor compared to inner-sphere chemical sorption
systems, where carbonaceous materials have shown to be effective. On and precipitation reactions that may occur with minerals and are usually
the other hand, methods that use adsorbents, such as activated C, ion not long-lived. The mechanism for the formation of outer- and
exchange resins, and zeolites, are costly and require expensive raw inner-sphere phosphate-metal complexes on biochar surfaces is pre­
materials (Kizito et al., 2015). Therefore, it is necessary to use affordable sented in Fig. 2. The mechanism of phosphate sorption by hydroxyl
adsorbents. In recent years, several studies have been conducted to functional groups on the biochar surface is described in Fig. 3.
investigate biochar performance as P adsorbents in aquatic systems. In Inner-sphere complex formation is accomplished by ligand exchange,
these studies, biochars had different P sorption capacities depending on but outer-sphere complex formation occurs through the formation of
the feedstock type and pyrolysis conditions. For instance, a study on the cationic bridges without ligand exchange (Shepherd et al., 2017).
phosphate sorption capacity of several biochar types produced at tem­ Factors affecting the sorption capacity of P are numerous, and in
peratures of 400–450 and 600–650 ◦ C showed that, as pyrolysis tem­ general, they are related to the properties of biochars and the solution
perature increased, the phosphate sorption capacity of oak wood and where the reaction is occurring. These factors include biochar’s specific
greenhouse waste biochars decreased, but the phosphate sorption ca­ surface area, the presence of metal oxides in biochar, and the pH of and
pacity of municipal waste and press cake biochars increased (Takaya the presence of other ions in the aqueous solution (Yin et al., 2017).
et al., 2016). A number of studies used biochars as adsorbents from P
removal from aquatic systems and then used those biochars as a
slow-release P fertilizer in a soil application. In this regard, biochars 5.3. Effects of biochar properties on phosphate sorption capacity
produced at 600 ◦ C from Mg-enriched tomato tissues are not only effi­
cient adsorbents for P removal from aqueous solutions but also a large 5.3.1. Biochar specific surface area
amount of P retained by the biochar was bioavailable, and the biochar Biochar can be effectively used to remove P from aqueous solutions
can be considered a slow-release P fertilizer (Yao et al., 2013a, 2013b, such as wastewater via the sorption process. The porous structure and
2013b). the high specific surface area of biochar allow sorption reactions and
transfer of P from an aqueous solution to the biochar surface. The spe­
cific surface area of biochar is a function of the feedstock type and py­
5.2. Mechanisms of P sorption by biochar rolysis temperature. As the pyrolysis temperature increases, the
biochar’s specific surface area increases due to the increased removal of
Phosphorus-biochar interactions are very complex, and our under­ volatiles from the feedstock, forming more pores. However, extremely
standing of these interactions is still limited. Some studies suggest that high temperatures lead to the disintegration of portions of biochar
biochar can adsorb P from aqueous solutions. Some other research, structure, blocking pores and reducing active sites. Several studies
however, shows that biochars having higher available P do not adsorb P investigated the impact of feedstock type on biochar’s specific surface
from aqueous solutions (Schneider and Haderlein, 2016). Phosphorus area. Under the same pyrolysis condition, biochars produced from wood
sorption by some biochars is accomplished through the interaction of P had higher specificity than those produced from rice husk (Kizito et al.,
with functional groups, involving metal cations, or through precipita­ 2015). Among biochars produced from peanut shells, oak, bamboo, and
tion reactions. Phosphate ions may also be adsorbed through ligand soybean stover under the same pyrolysis condition, biochars produced
exchange with hydroxyl and carboxyl groups on biochar surfaces from peanut shells had the highest specific surface area. Biochars pro­
(Shepherd et al., 2017). However, ligand exchange occurs only with duced from peanut shells, oak, bamboo, and soybean stover had a spe­
metal complexes. Phosphate ion species (PO3− 2−
4 , HPO4 , H2PO4 ) may

cific surface area of 328.96, 185.54, 110.52, and 247.09 m2 g− 1,

7
L. Ghodszad et al. Chemosphere 283 (2021) 131176

Fig. 3. Mechanism of phosphate sorption by hydroxyl groups on biochar: a) HPO2−


4 reacts with the water molecule to form H2PO4 and releases OH ion into the
− −

solution, b) H2PO−4 ion forming binuclear-bidentate bond with (Ox)FeOH groups on biochar surfaces, leading to the formation of phosphate-metal inner-sphere
complexes on biochar surfaces, and c) HPO2−
4 and H2PO4 ions release one and two OH ions, respectively, for the formation of inner-sphere complexes (modified
− −

from Shepherd et al., 2017).

respectively, with the highest phosphate sorption capacity in the peanut biochar metal oxides via the formation of strong chemical bonds. The
shell biochar (Jung et al., 2015a). Moreover, the phosphate sorption phosphate sorption capacity is correlated with the amount of calcium
capacity of marine macroalgae biochars was positively correlated with and magnesium present in biochars (Takaya et al., 2016). Phosphorus
the biochar’s specific surface area (Jung and Ahn, 2016). Hence, the ions can precipitate with free cations in the aqueous solution such as
specific surface area can be used as an index of biochar’s phosphate Ca2+ and Mg2+ that are released from biochar or co-precipitate with
sorption capacity. mixed mineral complexes (Al–Fe–Si–Ca) on biochar surfaces (Shepherd
et al., 2017). High pH, high Ca, Mg, Fe, and Al contents, the presence of
5.3.2. Metal oxides on biochar surface carbonates, especially calcite, explain the high P sorption capacity of a
Biochars may contain different types of metal oxides such as calcium pine wood-derived biochar (Bornø et al., 2018). Moreover, in Yao et al.
oxide, magnesium oxide, iron oxide, aluminum hydroxide, and (2013a, 2013b, 2011a, 2011b), the amount of magnesium in biochar
lanthanum oxide. These compounds can be effective in increasing the produced from sugar beet tailings and tomato treated with magnesium
phosphate sorption capacity by allowing phosphate in the solution to be nanocomposites was 10 and 8%, respectively; the magnesium resulted in
deposited on the biochar surface through weak hydrogen bonding with high phosphate sorption capacities (73 and 88.5%, respectively) in these
these metal oxides. Also, soluble phosphate can be precipitated with biochars. The researchers described the effective removal of phosphate

8
L. Ghodszad et al. Chemosphere 283 (2021) 131176

from aqueous solutions by these biochars as a result of the presence of biochar would more readily occur. The phosphate sorption capacity
magnesium nano-oxides, which were dispersed in large numbers on decreases at high pH values, as PO−4 3 is predominant at high pH and is
biochar surfaces. Many factors affect the high phosphate sorption ca­ hard to be adsorbed. The OH− and PO3− 4 ions also compete for sorption
pacity of those biochars that had high magnesium nano-oxide content. sites at high pH; this reduces the sorption of phosphate by biochar (Yin
For example, the surfaces of metal oxides are hydroxylated and et al., 2017).
depending on the pH of the aqueous solution, the electrical charge of
these surfaces can be positive or negative. The variation of charge of 5.4.2. Coexisting ions in aquatic systems
magnesium oxide on the biochar surface can be described as follows In natural aquatic systems, in addition to the target ion, other ions
(Stumm and Schindler, 1987): can affect the rate of sorption of the target ion onto biochar surfaces. In
most studies, the coexistence of phosphate and other ions in aqueous
SMgO − OH+
2 ⇌SMgO − OH⇌SMgO − O

solutions has a negative effect on phosphate sorption on biochar sur­
The pH at the point of zero-charge (pHpzc) of magnesium oxide is faces. For instance, the phosphate sorption capacity of biochar in syn­
about 12 (Yao et al., 2011a). If the pH of the aqueous solution is less than thetic solutions containing humic acidic, nitrate, chloride, bicarbonate,
the pHpzc, the magnesium oxide surface carries positive charges. and P decreased by up to 40% compared to solutions containing phos­
Otherwise, the magnesium oxide surface carries negative charges. phate only (Yao et al., 2013a, 2013b, 2013b). In another study, biochar
Magnesium oxide usually carries positive charges in most natural and produced from a mixture of several types of hardwoods absorbed 96 and
experimental aqueous solutions (most have pH < 12). Hence, phosphate 50% of phosphate in a pure phosphate solution and dairy effluent,
species are electrostatically adsorbed to the magnesium oxide surface respectively, after 24 h (Sarkhot et al., 2013). In situations where the
and form mono-, bi- and trinuclear complexes according to the following underlying mechanisms of adsorption of coexisting ions on biochar
reactions (Chernyakhovskii, 1985; Volceanov et al., 2002; Yao et al., differ from that of phosphate, the coexisting ions may have no
2011a): remarkable effect on the adsorption of phosphate. For example, high
0.12 > pH > 9.21 mononuclear complex SO2−
4 and NO3 concentrations had no remarkable effect on the adsorp­

tion of phosphate on the surface of biochars impregnated with ferric


SMgO − OH+ −
2 + H2 PO4 ⇌ ​ SMgO H2 PO4 + H2 O chloride (Yang et al., 2017); that was because phosphate adsorption
5.21 > pH > 10.67 binuclear complex occurred in oxides and hydroxides through ligand exchange, which
( ) formed inner-sphere complexes, while the adsorption of NO−3 and SO2− 4
was non-specifically adsorption and created the outer-sphere complexes,
2−
2SMgO − OH+2 + HPO4 ⇌ SMgO 2 HPO4 + 2H2 O

which would not interfere with the adsorption of phosphate. However,


10.67 > pH > 12 trinuclear complex
( ) the presence of CO2-3 could reduce the removal efficiency of phosphate by
3SMgO − OH+ 3−
2 + PO4 ⇌ SMgO 3 PO4 + 3H2 O about 20.6%, because CO2− 3 could compete intensely with phosphate for
the adsorption sites on biochar.
Also, some of the magnesium oxides can be dissolved and precipitate
Although biochar is recognized as an efficient and cost-effective
as Mg3 (PO4)2 and MgHPO4 through chemical reactions between phos­
adsorbent for environmental remediation, modification of biochar sur­
phate species and Mg2+ (Yao et al., 2011a).
faces can increase the P removal efficiency from aquatic systems
(Table 4). According to our review of the literature, most biochar
5.4. Effect of aquatic system properties on biochar’s phosphate sorption modification methods aimed to enhance biochar P sorption capacity by
capacity increasing biochar porosity, surface area and sorption sites (Jack et al.,
2019), increasing the abundance of oxygen-containing and organic
5.4.1. The pH in the aquatic system functional groups (Fang et al., 2014; Wang et al., 2020), enhancing Ca,
The pH of the aqueous solution can affect phosphate adsorption by Mg, Al, Fe and Mn nanoparticles/composites and generating nano-sized
changing the surface charge of the adsorbent and phosphate species in crystalline structures (Jung et al., 2015b), changing biochar surface
the solution. For instance, the pHpzc of biochars produced at tempera­ charge and zeta potential from negative to positive (Park et al., 2015;
tures above 400 ◦ C is higher than the pH of the aqueous solution, Yang et al., 2017), and increasing biochar anion exchange capacity
resulting in electrostatic adsorption between biochars and phosphates. (Mosa et al., 2018).
Biochar produced at 200 ◦ C, for example, had a low phosphate sorption
capacity, and the surface charge of this biochar was near zero when the 6. Conclusions and recommendations for future research
pH of the initial solution was 7 (Jung et al., 2015a). In addition, biochars
with a lower zeta potential might not be suitable for phosphate We conclude that biochar interacts with biotic and abiotic compo­
adsorption owing to their more negative surface charge (Trazzi et al., nents of the soil and directly and indirectly modifies soil P biochemical
2016). Overall, the zeta potential of biochars becomes more negative processes through altering P chemical forms, soil P sorption and
with increasing suspension pH, suggesting that the number of negative desorption capacities, microbial biomass, enzymatic activities, mycor­
charge increases and biochar becomes adverse to phosphate adsorption rhizal associations, and microbial production of metal-chelating organic
owing to the competition of OH− and phosphate for sorption sites and acids. The P bioavailability in biochar treated soils is controlled by soil
intensified the electrostatic repulsion between phosphate and the bio­ and biochar properties that are affected by feedstock type, pyrolysis
char (Yang et al., 2017). The fact that a high zeta potential means high condition, biochar application rate, and length of study. For improving
phosphate adsorption capacity has also been confirmed in the evalua­ soil fertility, obtaining biochar that is optimized for enhancing P
tion of phosphate removal by several types of biochars, such as bamboo, bioavailability for plants is desirable. Application of biochar in the soil
oak wood, soybean stover, peanut shell, and maize residue biochar increases soil CEC, changes soil pH, and releases organic ligands. These
(Jung et al., 2015a; Yin et al., 2017). The four different phosphate changes increase the adsorption of cations such as calcium, iron and
species, H3PO4, H2 PO−4 , and HPO2− 4 have pKa values of 2.12, 7.21, 12.67 aluminum, thereby reducing the adsorption of P and the precipitation of
in aqueous solutions (Jung et al., 2015a). Generally, the free energy of P in the soil, which ultimately leads to an increase in available P. In
sorption for H2PO−4 is lower than HPO2− 4 ; therefore, H2PO4 sorption by

9
L. Ghodszad et al. Chemosphere 283 (2021) 131176

Table 4
Phosphate sorption from aqueous solutions through various modified biochars.
Feedstock type Pyrolysis Treatment method Result Reference
temperature (◦ C)

Corn straw 800 Biochar decorated by NaLa(CO3)2 via Modified biochar had excellent monolayer PO3−4 uptake from water Qu et al.
one-pot hydrothermal method (330.9 mg g− 1) over a wide pH range of 3.0–8.0 (2020)
Corn stover 500 Corn stover treated with FeSO4 Modified biochar possessed 11–12 times more P sorption capacity Bakshi et al.
(2021)
Wheat straw 600 Biochar impregnated by solutions The maximum sorption capacity of biochar impregnated with the Zheng et al.
containing MgCl2, AlCl3, or both mixture of MgCl2, AlCl3 was 93.54 times the capacity of raw biochar (2020)
Ground corn 300, 450, and 600 The feedstock was dipped in MgCl2 Modification of biochar increased the maximum P adsorption from Fang et al.
solution 225 to 239 mg g− 1 (2014)
Raw corn straw 200, 300, 400, Raw corn straw was immersed in a ferrous The total P removal efficiency was >99% with the modified biochar Liu et al.
500, 600, 700 sulfate (FeSO4. 7H2O) solution and an effluent total P concentration was <0.02 mg L− 1 (2015)
Waste activated 550 Biochar was modified by chemical co- FeCl3-impregnated biochar had much better phosphate adsorption Yang et al.
sludge precipitation of Fe3+/Fe2+ or FeCl3 capacity (111.0 mg g− 1) in all as-prepared samples (2017)
impregnation
Wheat straw 400, 500 and 600 Wheat straw was soaked with bismuth Bismuth impregnated biochar produced at 500 ◦ C had high adsorption Zhu et al.
oxide solution and hydrochloric acid capacity for arsenic, P, and chromium (2016)
Wheat straw 600 Biochar loaded with CaO2 nanoparticles The nano-CaO2/biochar composite was efficient for phosphate Li et al.
removal in a wide range of initial pH and temperature, and the (2020b)
maximum adsorption capacity was 213.2 ± 13.6 mg g− 1
Sludge 550 Biochar peroxidized with nitric acid and The P adsorption capacity of the modified biochar increased ~ 40 Li et al.
impregnation with ferrous times (2020c)
Walnut shell 600 Impregnated with solutions with different The maximum sorption capacity of treated biochar increased three- Tao et al.
molar Fe2+ to Mg2+ ratios fold (2020a)
Cornstalk 550 Feedstock impregnated with MgCl2 Synthesized MgO-biochar with enough MgO active site had a Zhu et al.
maximum phosphate sorption capacity of 60.95 mg P g− 1 (2020)
Oak chips 500 Biochar loaded by lanthanum La-loaded biochar had greater adsorption ability due to decreased Wang et al.
surface negative charge and increased La-composites such as LaOOH, (2016)
LaONO3, and La(OH)3
Poplar chips 550 Chips were impregnated with AlCl3 Al-modified biochar had a greater P sorption capacity. Biochars with Yin et al.
20% Al had the highest PO3− 4 adsorption capacity (2018)
Moso bamboo 400, 500 and 600 Impregnation with MgCl2 Mg-laden biochar had higher phosphate adsorption capacities. The Jiang et al.
(Phyllostachys maximum phosphate adsorption amount was 344, 357, and 370 mg (2018)
pubescens) g− 1 for biochar-Mg-400, biochar-Mg-500, and biochar-Mg-600,
respectively
Sugarcane harvest 550 The residue was impregnated with MgCl2 Mg introduction enhanced biochar’s affinity for PO3-
4 Li et al.
residue (2017a)
Bamboo 300, 450 and 600 Biochar modified by polyethyleneimine The P adsorption capacity of PEI-modified biochar was enhanced with Li et al.
(PEI) increasing pyrolysis temperature from 300 to 450 ◦ C, but not with (2019b)
600 ◦ C
Bamboo 600 Biochar functionalized with Mg–Al and Composite with 40% Mg–Al LDH had the highest PO3− 4 adsorption Wan et al.
Mg–Fe (3:1) layered double hydroxides with >95% PO3− 4 saturation achieved within 1 h (2017)
(LDH)
Cotton stalk 350 Granulation of biochar powder followed Both granulation and ferric oxide loading increased the surface area Ren et al.
by impregnation with ferric oxides and the adsorption capacity (from 0 to 0.963 mg P g− 1) (2015)
Carrot residue 200, 300, 400, Carrot powder was added to a MgCl2 Biochar produced at 400 ◦ C and impregnated with Mg had higher Pinto et al.
500 and 600 solution efficiencies for P removal from aqueous solutions (2019)
Hickory wood chips 600 The feedstock was immersed in an AlCl3 Engineered biochar effectively removed P from a secondary treated Zheng et al.
solution wastewater (2019)

addition, biochar can directly and indirectly affect microbial and biochar’s P sorption capacities. To improve the efficiency of biochar for
enzymatic activities and change P availability. In this regard, biochar P capture from aqueous solutions, factors such as biochar chemical
increases soil water retention, internal surfaces, and microbial popula­ composition, biochar surface physical properties, pH, and coexisting
tion and ultimately increases the activities of enzymes, including ions should be considered.
phosphatases. In addition, biochar can be a source of P and increase P There are research gaps that need to be addressed in future research
supply. On the other hand, P precipitation with minerals in the biochar regarding biochar application for P management: 1.) despite numerous
and soil may occur. studies on soil P bioavailability changes upon biochar addition, the
In addition, biochar has the capacity to remove and recover P from underlying mechanisms remain poorly understood and should be further
wastewater. Physicochemical reactions govern the P sorption capacity studied; 2.) the economic feasibility of biochar application should be
from wastewater in various types of biochars. The P sorption from water part of future research. If biochar application is not economically viable,
systems by biochar depends on the feedstock type and pyrolysis condi­ such applications would not be sustainable. Strategies for reducing
tion used for biochar production. Several mechanisms have been pro­ biochar losses and cost of applications and enhancing its use efficiency
posed for P removal from water systems. Adsorption of P is through the need to be developed; 3.) the longevity of biochar benefits on P cycling
interaction of phosphate with biochar surface functional groups such as and related soil fertility needs to be addressed in long-term field studies;
carboxyl and hydroxyl through the formation of hydrogen bonds and 4.) interactions among biochar, soil, microorganisms and plants in
electrostatic interactions. Biochar can also adsorb P by ligand exchange biochar-treated soils are complicated, and mechanistic understanding of
with metal complexes or cationic bridges. Properties such as pH and those interactions is still lacking. More research to understand those
coexisting ions in water systems also affect biochar’s P sorption capac­ interactions are urgently needed so that biochar applications are better
ity. Increasing the pH and CO2− 3 concentration in the water system re­ guided; 5.) the relative contribution of desorbed P from biochar or soil to
duces the biochar P sorption capacity. Various modification methods soil P availability could be studied using the 32P isotope technique. The
32
can be used to alter biochar surface properties, which can enhance P isotope technique should help us to understand the mechanisms of P

10
L. Ghodszad et al. Chemosphere 283 (2021) 131176

sorption/desorption by soil constituents in the presence of biochar; 6.) Dai, Y., Wang, W., Lu, L., Yan, L., Yu, D., 2020. Utilization of biochar for the removal of
nitrogen and phosphorus. J. Clean. Prod. 257, 120573. https://doi.org/10.1016/j.
more research is required to develop low-cost modification methods to
jclepro.2020.120573.
increase the P sorption capacity of biochars and reuse such biochars as a Dari, B., Nair, V.D., Harris, W.G., Nair, P.K.R., Sollenberger, L., Mylavarapu, R., 2016.
slow-release P fertilizer, and 7.) systematic studies on the potential Relative influence of soil- vs. biochar properties on soil phosphorus retention.
ecological risk of applied biochar in soil or water systems should be Geoderma 280, 82–87. https://doi.org/10.1016/J.GEODERMA.2016.06.018.
DeLuca, T.H., MacKenzie, M.D., Gundale, M.J., 2009. Biochar effects on soil nutrient
conducted. For instance, the toxicity (e.g., heavy metals) and salinity transformations. In: Lehmann, J., Joseph, S. (Eds.), Biochar for Environmental
risk of biochar leachate of some biochars can have adverse effects on the Management: Science, Technology and Implementation. Earthscan, London, United
environment. Kingdom, pp. 251–270.
DeLuca, T.H., Gundale, M.J., MacKenzie, M.D., Jones, D.L., 2015. Biochar effects on soil
nutrient transformations. In: Lehmann, J., Joseph, S. (Eds.), Biochar for
Environmental Management. Science, Technology and Implementation. Taylor &
Declaration of competing interest Francis, pp. 421–454.
Drizo, A., Frost, C.A., Grace, J., Smith, K.A., 1999. Physico-chemical screening of
phosphate-removing substrates for use in constructed wetland systems. Water Res.
The authors declare that they have no known competing financial 33, 3595–3602. https://doi.org/10.1016/S0043-1354(99)00082-2.
interests or personal relationships that could have appeared to influence Du, Z.Y., Wang, Q.H., Liu, F.C., Ma, H.L., Ma, B.Y., Malhi, S.S., 2013. Movement of
the work reported in this paper. phosphorus in a calcareous soil as affected by humic acid. Pedosphere 23, 229–235.
https://doi.org/10.1016/S1002-0160(13)60011-9.
Dume, B., Tessema, D.A., Regassa, A., Berecha, G., 2017. Effects of biochar on
Acknowledgements phosphorus sorption and desorption in acidic and calcareous soils. Civ. Environ. Res.
9, 10–20.
Eduah, J.O., Nartey, E.K., Abekoe, M.K., Breuning-Madsen, H., Andersen, M.N., 2019.
We thank the editor and two anonymous reviewers whose Phosphorus retention and availability in three contrasting soils amended with rice
constructive comments improved the quality of an earlier version of this husk and corn cob biochar at varying pyrolysis temperatures. Geoderma 341, 10–17.
manuscript. https://doi.org/10.1016/j.geoderma.2019.01.016.
Efthymiou, A., Grønlund, M., Müller-Stöver, D.S., Jakobsen, I., 2018. Augmentation of
the phosphorus fertilizer value of biochar by inoculation of wheat with selected
References Penicillium strains. Soil Biol. Biochem. 116, 139–147. https://doi.org/10.1016/j.
soilbio.2017.10.006.
El-Naggar, A., Lee, S.S., Rinklebe, J., Farooq, M., Song, H., Sarmah, A.K., Zimmerman, A.
Al Marzooqi, F., Yousef, L.F., 2017. Biological response of a sandy soil treated with
R., Ahmad, M., Shaheen, S.M., Ok, Y.S., 2019. Biochar application to low fertility
biochar derived from a halophyte (Salicornia bigelovii). Appl. Soil Ecol. 114, 9–15.
soils: a review of current status, and future prospects. Geoderma 337, 536–554.
https://doi.org/10.1016/j.apsoil.2017.02.012.
https://doi.org/10.1016/j.geoderma.2018.09.034.
Amarakoon, I., Zvomuya, F., Motaung, M.L., 2019. Temperature-dependency of
Enders, A., Hanley, K., Whitman, T., Joseph, S., Lehmann, J., 2012. Characterization of
phosphorus sorption by goethites and tropical soils amended with woodchip biochar.
biochars to evaluate recalcitrance and agronomic performance. Bioresour. Technol.
Agrosyst. Geosci. Environ. 2, 1–6. https://doi.org/10.2134/age2018.12.0067.
114, 644–653. https://doi.org/10.1016/j.biortech.2012.03.022.
Antoniadis, V., Koliniati, R., Efstratiou, E., Golia, E., Petropoulos, S., 2016. Effect of soils
Enders, A., Sohi, S., Lehmann, J., Sing, B., 2017. Total elemental analysis of metals and
with varying degree of weathering and pH values on phosphorus sorption. Catena
nutrients in biochars. In: Singh, B., Camps-Arbestain, M., Lehmann, J. (Eds.),
139, 214–219. https://doi.org/10.1016/j.catena.2016.01.008.
Biochar: A Guide to Analytical Methods. CSIRO Publishing, CRC Press, pp. 95–108.
Bakshi, S., Laird, D.A., Smith, R.G., Brown, R.C., 2021. Capture and release of
Fang, C., Zhang, T., Li, P., Jiang, R.F., Wang, Y.C., 2014. Application of magnesium
orthophosphate by Fe-modified biochars: mechanisms and environmental
modified corn biochar for phosphorus removal and recovery from swine wastewater.
applications. ACS Sustain. Chem. Eng.658–668 ,(2) 9 . https://doi.org/10.1021/
Int. J. Environ. Res. Publ. Health 11, 9217–9237. https://doi.org/10.3390/
acssuschemeng.0c06108.
ijerph110909217.
Bhaduri, D., Saha, A., Desai, D., Meena, H.N., 2016. Restoration of carbon and microbial
Gao, S., DeLuca, T., 2016. Influence of biochar on soil nutrient transformations, nutrient
activity in salt-induced soil by application of peanut shell biochar during short-term
leaching, and crop yield. Adv. Plants Agric. Res. 4, 00150 https://doi.org/10.15406/
incubation study. Chemosphere 148, 86–98. https://doi.org/10.1016/j.
apar.2016.04.00150.
chemosphere.2015.12.130.
Gao, S., DeLuca, T.H., Cleveland, C.C., 2019. Biochar additions alter phosphorus and
Böhme, L., Langer, U., Böhme, F., 2005. Microbial biomass, enzyme activities and
nitrogen availability in agricultural ecosystems: a meta-analysis. Sci. Total Environ.
microbial community structure in two European long-term field experiments. Agric.
654, 463–472. https://doi.org/10.1016/j.scitotenv.2018.11.124.
Ecosyst. Environ. 109, 141–152. https://doi.org/10.1016/j.agee.2005.01.017.
Gul, S., Whalen, J.K., 2016. Biochemical cycling of nitrogen and phosphorus in biochar-
Bornø, M.L., Müller-Stöver, D.S., Liu, F., 2018. Contrasting effects of biochar on
amended soils. Soil Biol. Biochem. 103, 1–15. https://doi.org/10.1016/j.
phosphorus dynamics and bioavailability in different soil types. Sci. Total Environ.
soilbio.2016.08.001.
627, 963–974. https://doi.org/10.1016/J.SCITOTENV.2018.01.283.
Gul, S., Whalen, J.K., Thomas, B.W., Sachdeva, V., Deng, H., 2015. Physico-chemical
Bruun, E.W., Ambus, P., Egsgaard, H., Hauggaard-Nielsen, H., 2012. Effects of slow and
properties and microbial responses in biochar-amended soils: mechanisms and future
fast pyrolysis biochar on soil C and N turnover dynamics. Soil Biol. Biochem. 46,
directions. Agric. Ecosyst. Environ. 206, 46–59. https://doi.org/10.1016/j.
73–79. https://doi.org/10.1016/j.soilbio.2011.11.019.
agee.2015.03.015.
Chandra, S., Bhattacharya, J., 2019. Influence of temperature and duration of pyrolysis
Gunes, A., Inal, A., Taskin, M.B., Sahin, O., Kaya, E.C., Atakol, A., 2014. Effect of
on the property heterogeneity of rice straw biochar and optimization of pyrolysis
phosphorus-enriched biochar and poultry manure on growth and mineral
conditions for its application in soils. J. Clean. Prod. 215, 1123–1139. https://doi.
composition of lettuce (Lactuca sativa L. cv.) grown in alkaline soil. Soil Use Manag.
org/10.1016/J.JCLEPRO.2019.01.079.
30, 182–188. https://doi.org/10.1111/sum.12114.
Chen, J., Liu, X., Zheng, J., Zhang, B., Lu, H., Chi, Z., Pan, G., Li, L., Zheng, J., Zhang, X.,
Hale, S.E., Lehmann, J., Rutherford, D., Zimmerman, A.R., Bachmann, R.T.,
Wang, J., Yu, X., 2013. Biochar soil amendment increased bacterial but decreased
Shitumbanuma, V., O’Toole, A., Sundqvist, K.L., Arp, H.P.H., Cornelissen, G., 2012.
fungal gene abundance with shifts in community structure in a slightly acid rice
Quantifying the total and bioavailable polycyclic aromatic hydrocarbons and dioxins
paddy from Southwest China. Appl. Soil Ecol. 71, 33–44. https://doi.org/10.1016/j.
in biochars. Environ. Sci. Technol. 46, 2830–2838. https://doi.org/10.1021/
apsoil.2013.05.003.
es203984k.
Chen, M., Alim, N., Zhang, Y., Xu, N., Cao, X., 2018. Contrasting effects of biochar
Hammer, E.C., Balogh-Brunstad, Z., Jakobsen, I., Olsson, P.A., Stipp, S.L., Rillig, M.C.,
nanoparticles on the retention and transport of phosphorus in acidic and alkaline
2014. A mycorrhizal fungus grows on biochar and captures phosphorus from its
soils. Environ. Pollut. 239, 562–570. https://doi.org/10.1016/j.envpol.2018.04.050.
surfaces. Soil Biol. Biochem. 77, 252–260. https://doi.org/10.1016/j.
Chen, Q., Qin, J., Cheng, Z., Huang, L., Sun, P., Chen, L., Shen, G., 2018. Synthesis of a
soilbio.2014.06.012.
stable magnesium-impregnated biochar and its reduction of phosphorus leaching
Hammer, E.C., Forstreuter, M., Rillig, M.C., Kohler, J., 2015. Biochar increases
from soil. Chemosphere 199, 402–408. https://doi.org/10.1016/j.
arbuscular mycorrhizal plant growth enhancement and ameliorates salinity stress.
chemosphere.2018.02.058.
Appl. Soil Ecol. 96, 114–121. https://doi.org/10.1016/j.apsoil.2015.07.014.
Chernyakhovskii, V.A., 1985. Technology of unfired periclase-spinel parts with a
Hilber, I., Blum, F., Leifeld, J., Schmidt, H.P., Bucheli, T.D., 2012. Quantitative
phosphate binder. Refractories 26, 41–44. https://doi.org/10.1007/BF01398613.
determination of PAHs in biochar: a prerequisite to ensure its quality and safe
Chintala, R., Schumacher, T.E., Mcdonald, L.M., Clay, D.E., Malo, D.D., Papiernik, S.K.,
application. J. Agric. Food Chem. 60, 3042–3050. https://doi.org/10.1021/
Clay, S.A., Julson, J.L., 2014. Phosphorus sorption and availability from biochars
jf205278v.
and soil/biochar mixtures. Clean 42, 626–634. https://doi.org/10.1002/
Hong, C., Lu, S., 2018. Does biochar affect the availability and chemical fractionation of
clen.201300089.
phosphate in soils? Environ. Sci. Pollut. Res. 25, 8725–8734. https://doi.org/
Cui, L., Yin, C., Chen, T., Quan, G., Ippolito, J.A., Liu, B., Yan, J., Ding, C., Hussain, Q.,
10.1007/s11356-018-1219-8.
Umer, M., 2019. Remediation of organic halogen- contaminated wetland soils using
Jack, J., Huggins, T.M., Huang, Y., Fang, Y., Ren, Z.J., 2019. Production of magnetic
biochar. Sci. Total Environ. 696, 134087. https://doi.org/10.1016/j.
biochar from waste-derived fungal biomass for phosphorus removal and recovery.
scitotenv.2019.134087.
J. Clean. Prod. 224, 100–106. https://doi.org/10.1016/j.jclepro.2019.03.120.
Dai, Z., Zhang, X., Tang, C., Muhammad, N., Wu, J., Brookes, P.C., Xu, J., 2017. Potential
Jiang, D., Chu, B., Amano, Y., Machida, M., 2018. Removal and recovery of phosphate
role of biochars in decreasing soil acidification - a critical review. Sci. Total Environ.
from water by Mg-laden biochar: batch and column studies. Colloid. Surf. A -
581, 601–611. https://doi.org/10.1016/j.scitotenv.2016.12.169.

11
L. Ghodszad et al. Chemosphere 283 (2021) 131176

Physicochem. Eng. Asp. 558, 429–437. https://doi.org/10.1016/j. soils. Chemosphere 235, 32–39. https://doi.org/10.1016/j.
colsurfa.2018.09.016. chemosphere.2019.06.160.
Jin, Y., Liang, X., He, M., Liu, Y., Tian, G., Shi, J., 2016. Manure biochar influence upon Martínez, C.M.J., España, A.J.C., Díaz, V.J.J., 2017. Effect of Eucalyptus globullus
soil properties, phosphorus distribution and phosphatase activities: a microcosm biochar addition on the availability of phosphorus in acidic soil. Agron. Colomb. 35,
incubation study. Chemosphere 142, 128–135. https://doi.org/10.1016/j. 75–81. https://doi.org/10.15446/agron.colomb.v35n1.58671.
chemosphere.2015.07.015. Masto, E.R., Kumar, S., Rout, T., Sarkar, P., George, J., Ram, L., 2013. Biochar from water
Jung, K.W., Ahn, K.H., 2016. Fabrication of porosity-enhanced MgO/biochar for removal hyacinth (Eichornia crassipes) and its impact on soil biological activity. Catena 111,
of phosphate from aqueous solution: application of a novel combined 64–71. https://doi.org/10.1016/j.catena.2013.06.025.
electrochemical modification method. Bioresour. Technol. 200, 1029–1032. https:// Mendes, G., Zafra, D.L., Vassilev, N.B., Silva, I.R., Ribeiro, J.I., Costaa, M.D., 2014.
doi.org/10.1016/j.biortech.2015.10.008. Biochar enhances aspergillus Niger rock phosphate solubilization by increasing
Jung, K.W., Hwang, M.J., Ahn, K.H., Ok, Y.S., 2015a. Kinetic study on phosphate organic acid production and alleviating fluoride toxicity. Appl. Environ. Microbiol.
removal from aqueous solution by biochar derived from peanut shell as renewable 80, 3081–3085. https://doi.org/10.1128/AEM.00241-14.
adsorptive media. Int. J. Environ. Sci. Technol. 12, 3363–3372. https://doi.org/ Mendes, J.D., Chaves, L.H.D., Chaves, I.D.D., Silva, F.D.A.S.E., Fernandes, J.D., 2015.
10.1007/s13762-015-0766-5. Using poultry litter biochar and rock dust MB-4 on release available phosphorus to
Jung, K.W., Hwang, M.J., Jeong, T.U., Ahn, K.H., 2015b. A novel approach for soils. Agric. Sci. 1367–1374. https://doi.org/10.4236/as.2015.611131, 06.
preparation of modified-biochar derived from marine macroalgae: dual purpose Moinet, G.Y.K., Hunt, J.E., Kirschbaum, M.U.F., Morcom, C.P., Midwood, A.J.,
electro-modification for improvement of surface area and metal impregnation. Millard, P., 2018. The temperature sensitivity of soil organic matter decomposition is
Bioresour. Technol. 191, 342–345. https://doi.org/10.1016/j.biortech.2015.05.052. constrained by microbial access to substrates. Soil Biol. Biochem. 116, 333–339.
Kambo, H.S., Dutta, A., 2015. A comparative review of biochar and hydrochar in terms of https://doi.org/10.1016/j.soilbio.2017.10.031.
production, physico-chemical properties and applications. Renew. Sustain. Energy Morales, M.M., Comerford, N., Guerrini, I.A., Falcão, N.P.S., Reeves, J.B., 2013. Sorption
Rev. 45, 359–378. https://doi.org/10.1016/j.rser.2015.01.050. and desorption of phosphate on biochar and biochar-soil mixtures. Soil Use Manag.
Khadem, A., Raiesi, F., 2019. Response of soil alkaline phosphatase to biochar 29, 306–314. https://doi.org/10.1111/sum.12047.
amendments: changes in kinetic and thermodynamic characteristics. Geoderma 337, Mosa, A., El-Ghamry, A., Tolba, M., 2018. Functionalized biochar derived from heavy
44–54. https://doi.org/10.1016/j.geoderma.2018.09.001. metal rich feedstock: phosphate recovery and reusing the exhausted biochar as an
Kizito, S., Wu, S., Kirui, W.K., Lei, M., Lu, Q., Bah, H., Dong, R., 2015. Evaluation of slow enriched soil amendment. Chemosphere 198, 351–363. https://doi.org/10.1016/j.
pyrolyzed wood and rice husks biochar for adsorption of ammonium nitrogen from chemosphere.2018.01.113.
piggery manure anaerobic digestate slurry. Sci. Total Environ. 505, 102–112. Mukherjee, A., Zimmerman, A.R., Harris, W., 2011. Surface chemistry variations among
https://doi.org/10.1016/j.scitotenv.2014.09.096. a series of laboratory-produced biochars. Geoderma 163, 247–255. https://doi.org/
Kumari, K.G.I.D., Moldrup, P., Paradelo, M., Elsgaard, L., Hauggaard-Nielsen, H., 10.1016/j.geoderma.2011.04.021.
Jonge, L.W.D., 2014. Effects of biochar on air and water permeability and colloid Mukherjee, S., Mavi, M.S., Singh, J., Singh, B.P., 2019. Rice-residue biochar influences
and phosphorus leaching in soils from a natural calcium carbonate gradient. phosphorus availability in soil with contrasting P status. Arch. Agron Soil Sci. 66 (6),
J. Environ. Qual. 43, 647–657. https://doi.org/10.2134/jeq2013.08.0334. 778–791. https://doi.org/10.1080/03650340.2019.1639153.
Laghari, M., Mirjat, M.S., Hu, Z., Fazal, S., Xiao, B., Hu, M., Chen, Z., Guo, D., 2015. Mukome, F.N.D., Zhang, X., Silva, L.C.R., Six, J., Parikh, S.J., 2013. Use of chemical and
Effects of biochar application rate on sandy desert soil properties and sorghum physical characteristics to investigate trends in biochar feedstocks. J. Agric. Food
growth. Catena 135, 313–320. https://doi.org/10.1016/j.catena.2015.08.013. Chem. 61, 2196–2204. https://doi.org/10.1021/jf3049142.
Laird, D., Fleming, P., Wang, B., Horton, R., Karlen, D., 2010. Biochar impact on nutrient Narula, N., Kumar, V., Behl, R.K., Deubel, A., Gransee, A., Merbach, W., 2000. Effect of P-
leaching from a Midwestern agricultural soil. Geoderma 158, 436–442. https://doi. solubilizing Azotobacter chroococcum on N, P, K uptake in P-responsive wheat
org/10.1016/j.geoderma.2010.05.012. genotypes grown under greenhouse conditions. J. Plant Nutr. Soil Sci. 163, 393–398.
Lee, J.W., Kidder, M., Evans, B.R., Paik, S., Buchanan, A.C., Garten, C.T., Brown, R.C., https://doi.org/10.1002/1522-2624, 200008)163:4<393::AID-JPLN393>3.0.CO;2-
2010. Characterization of biochars produced from cornstovers for soil amendment. W.
Environ. Sci. Technol. 44, 7970–7974. https://doi.org/10.1021/es101337x. Nigam, N., Yadav, V., Mishra, D., Karak, T., Khare, P., 2019. Biochar amendment alters
Li, X., Shen, Q., Zhang, D., Mei, X., Ran, W., Xu, Y., Yu, G., 2013. Functional groups the relation between the Pb distribution and biological activities in soil. Int. J.
determine biochar properties (pH and EC) as studied by two-dimensional 13C NMR Environ. Sci. Technol. 16, 8595–8606. https://doi.org/10.1007/s13762-019-02257-
correlation spectroscopy. PloS One 8. https://doi.org/10.1371/journal. y.
pone.0065949. Nobaharan, K., Bagheri Novair, S., Asgari Lajayer, B., van Hullebusch, E.D., 2021.
Li, H., Dong, X., Silva, E.B.D., Oliveira, L.M.D., Chen, Y., Ma, L.Q., 2017. Mechanisms of Phosphorus removal from wastewater: the potential use of biochar and the key
metal sorption by biochars: biochar characteristics and modifications. Chemosphere controlling factors. Water 13, 517. https://doi.org/10.3390/w13040517.
178, 466–478. https://doi.org/10.1016/j.chemosphere.2017.03.072. Ogawa, M., Okimori, Y., 2010. Pioneering works in biochar research, Japan. Aust. J. Soil
Li, R., Wang, J.J., Zhou, B., Zhang, Z., Liu, S., Lei, S., Xiao, R., 2017. Simultaneous Res. 48, 489–500. https://doi.org/10.1071/SR10006.
capture removal of phosphate, ammonium and organic substances by MgO Ok, Y., Uchimiya, S., Chang, S., Bolan, N., 2015. Biochar: Production, Characterization,
impregnated biochar and its potential use in swine wastewater treatment. J. Clean. and Applications. CRC Press, New York, USA.
Prod. 147, 96–107. https://doi.org/10.1016/j.jclepro.2017.01.069. Oladele, S.O., 2019. Effect of biochar amendment on soil enzymatic activities,
Li, S., Harris, S., Anandhi, A., Chen, G., 2019. Predicting biochar properties and functions carboxylate secretions and upland rice performance in a sandy clay loam Alfisol of
based on feedstock and pyrolysis temperature: a review and data syntheses. J. Clean. Southwest Nigeria. Sci. African 4, e00107. https://doi.org/10.1016/j.sciaf.2019.
Prod. 215, 890–902. https://doi.org/10.1016/j.jclepro.2019.01.106. e00107.
Li, T., Tong, Z., Gao, B., Li, Y.C., Smyth, A., Bayabil, H.K., 2019. Polyethyleneimine- Park, J.H., Ok, Y.S., Kim, S.H., Cho, J.S., Heo, J.S., Delaune, R.D., Seo, D.C., 2015.
modified biochar for enhanced phosphate adsorption. Environ. Sci. Pollut. Res. 27, Evaluation of phosphorus adsorption capacity of sesame straw biochar on aqueous
7420–7429. https://doi.org/10.1007/s11356-019-07053-2. solution: influence of activation methods and pyrolysis temperatures. Environ.
Li, H., Li, Y., Xu, Y., Lu, X., 2020a. Biochar phosphorus fertilizer effects on soil Geochem. Health 37, 969–983. https://doi.org/10.1007/s10653-015-9709-9.
phosphorus availability. Chemosphere 244, 125471. https://doi.org/10.1016/j. Peng, C., Li, Q., Zhang, Z., Wu, Z., Song, X., Zhou, G., Song, X., 2019. Biochar amendment
chemosphere.2019.125471. changes the effects of nitrogen deposition on soil enzyme activities in a Moso
Li, X., Xie, Y., Jiang, F., Wang, B., Hu, Q., Tang, Y., Luo, T., Wu, T., 2020b. Enhanced bamboo plantation. J. For. Res. 24, 275–284. https://doi.org/10.1080/
phosphate removal from aqueous solution using resourceable nano-CaO2/BC 13416979.2019.1646970.
composite: behaviors and mechanisms. Sci. Total Environ. 709, 136123. https://doi. Pinto, M.C.E., Silva, D.D.D., Gomes, A.L.A., Santos, R.M.M.D., Couto, R.A.A.D.,
org/10.1016/j.scitotenv.2019.136123. Novais, R.F.D., Constantino, V.R.L., Tronto, J., Pinto, F.G., 2019. Biochar from carrot
Li, Z., Liu, X., Wang, Y., 2020c. Modification of sludge-based biochar and its application residues chemically modified with magnesium for removing phosphorus from
to phosphorus adsorption from aqueous solution. J. Mater. Cycles Waste Manag. 22, aqueous solution. J. Clean. Prod. 222, 36–46. https://doi.org/10.1016/j.
123–132. https://doi.org/10.1007/s10163-019-00921-6. jclepro.2019.03.012.
Liu, F., Zuo, J., Chi, T., Wang, P., Yang, B., 2015. Removing phosphorus from aqueous Qian, T., Yang, Q., Jun, D.C.F., Dong, F., Zhou, Y., 2019. Transformation of phosphorus
solutions by using iron-modified corn straw biochar. Front. Environ. Sci. Eng. 9, in sewage sludge biochar mediated by a phosphate-solubilizing microorganism.
1066–1075. https://doi.org/10.1007/s11783-015-0769-y. Chem. Eng. J. 359, 1573–1580. https://doi.org/10.1016/j.cej.2018.11.015.
Liu, S., Meng, J., Jiang, L., Yang, X., Lan, Y., Cheng, X., Chen, W., 2017a. Rice husk Qu, J., Akindolie, M.S., Feng, Y., Jiang, Z., Zhang, G., Jiang, Q., Fengxia, D., Cao, B.,
biochar impacts soil phosphorous availability, phosphatase activities and bacterial Zhang, Y., 2020. One-pot hydrothermal synthesis of NaLa (CO3)2 decorated magnetic
community characteristics in three different soil types. Appl. Soil Ecol. 116, 12–22. biochar for efficient phosphate removal from water: kinetics, isotherms,
https://doi.org/10.1016/j.apsoil.2017.03.020. thermodynamics, mechanisms and reusability exploration. Chem. Eng. J. 394,
Liu, C., Liu, F., Ravnskov, S., Rubæk, G.H., Sun, Z., Andersen, M.N., 2017b. Impact of 124915. https://doi.org/10.1016/j.cej.2020.124915.
wood biochar and its interactions with mycorrhizal fungi, phosphorus fertilization Rafiq, M.K., Bachmann, R.T., Rafiq, M.T., Shang, Z., Joseph, S., Long, R.L., 2016.
and irrigation strategies on potato growth. J. Agron. Crop Sci. 203, 131–145. Influence of pyrolysis temperature on physico-chemical properties of corn stover
https://doi.org/10.1111/jac.12185 . (Zea mays L.) biochar and feasibility for carbon capture and energy balance. PloS
Liu, Y., Zhu, Z.Q., He, X.S., Yang, C., Du, Y.Q., Huang, Y.D., Su, P., Wang, S., Zheng, X.X., One 11, e0156894. https://doi.org/10.1371/journal.pone.0156894.
Xue, Y.J., 2018. Mechanisms of rice straw biochar effects on phosphorus sorption Ren, J., Li, N., Li, L., An, J.K., Zhao, L., Ren, N.Q., 2015. Granulation and ferric oxides
characteristics of acid upland red soils. Chemosphere 207, 267–277. https://doi.org/ loading enable biochar derived from cotton stalk to remove phosphate from water.
10.1016/j.chemosphere.2018.05.086. Bioresour. Technol. 178, 119–125. https://doi.org/10.1016/j.biortech.2014.09.071.
Liu, M., Che, Y., Wang, L., Zhao, Z., Zhang, Y., Wei, L., Xiao, Y., 2019. Rice straw biochar Riddle, M., Bergström, L., Schmieder, F., Lundberg, D., Condron, L., Cederlund, H., 2019.
and phosphorus inputs have more positive effects on the yield and nutrient uptake of Impact of biochar coated with magnesium (hydr)oxide on phosphorus leaching from
Lolium multiflorum than arbuscular mycorrhizal fungi in acidic Cd-contaminated

12
L. Ghodszad et al. Chemosphere 283 (2021) 131176

organic and mineral soils. J. Soils Sediments 19, 1875–1889. https://doi.org/ activities and fungal communities in replant disease soil. Sci. Hortic. (Amst.) 256,
10.1007/s11368-018-2197-7. 108641. https://doi.org/10.1016/j.scienta.2019.108641.
Ronsse, F., Hecke, S.V., Dickinson, D., Prins, W., 2013. Production and characterization Wang, B., Lian, G., Lee, X., Gao, B., Li, L., Liu, T., Zhang, X., Zheng, Y., 2020.
of slow pyrolysis biochar: influence of feedstock type and pyrolysis conditions. GCB Phosphogypsum as a novel modifier for distillers grains biochar removal of
Bioenergy 5, 104–115. https://doi.org/10.1111/gcbb.12018. phosphate from water. Chemosphere 238, 124684. https://doi.org/10.1016/j.
Sarkhot, D.V., Ghezzehei, T.A., Berhe, A.A., 2013. Effectiveness of biochar for sorption of chemosphere.2019.124684.
ammonium and phosphate from dairy effluent. J. Environ. Qual. 42, 1545–1554. Warnock, D.D., Mummey, D.L., McBride, B., Major, J., Lehmann, J., Rillig, M.C., 2010.
https://doi.org/10.2134/jeq2012.0482. Influences of non-herbaceous biochar on arbuscular mycorrhizal fungal abundances
Schneider, F., Haderlein, S.B., 2016. Potential effects of biochar on the availability of in roots and soils: results from growth-chamber and field experiments. Appl. Soil
phosphorus - mechanistic insights. Geoderma 277, 83–90. https://doi.org/10.1016/ Ecol. 46, 450–456. https://doi.org/10.1016/j.apsoil.2010.09.002.
j.geoderma.2016.05.007. Wei, S., Zhu, M., Fan, X., Song, J., Peng, P., Li, K., Jia, W., Song, H., 2019. Influence of
Shepherd, J.G., Joseph, S., Sohi, S.P., Heal, K.V., 2017. Biochar and enhanced phosphate pyrolysis temperature and feedstock on carbon fractions of biochar produced from
capture: mapping mechanisms to functional properties. Chemosphere 179, 57–74. pyrolysis of rice straw, pine wood, pig manure and sewage sludge. Chemosphere
https://doi.org/10.1016/j.chemosphere.2017.02.123. 218, 624–631. https://doi.org/10.1016/j.chemosphere.2018.11.177.
Sial, T.A., Khan, M.N., Lan, Z., Kumbhar, F., Ying, Z., Zhang, J., Sun, D., Li, X., 2019. Winsley, P., 2007. Biochar and bioenergy production for climate change mitigation. N. Z.
Contrasting effects of banana peels waste and its biochar on greenhouse gas Sci. Rev. 64, 5–10.
emissions and soil biochemical properties. Process Saf. Environ. Protect. 122, Xu, G., Sun, J.N., Shao, H.B., Chang, S.X., 2014. Biochar had effects on phosphorus
366–377. https://doi.org/10.1016/j.psep.2018.10.030. sorption and desorption in three soils with differing acidity. Ecol. Eng. 62, 54–60.
Siddiqui, A.R., Nazeer, S., Piracha, M.A., Saleem, M.M., Siddiqi, I., Shahzad, S.M., https://doi.org/10.1016/j.ecoleng.2013.10.027.
Sarwar, G., 2016. The production of biochar and its possible effects on soil properties Xu, M., Gao, P., Yang, Z., Su, L., Wu, J., Yang, G., Zhang, X., Ma, J., Peng, H., Xiao, Y.,
and phosphate solubilizing bacteria. J. Appl. Agric. Biotechnol. 1, 27–40. . 2019. Biochar impacts on phosphorus cycling in rice ecosystem. Chemosphere 225,
Silva, L.G., Andrade, C.A.D., Bettiol, W., 2020. Biochar amendment increases soil 311–319. https://doi.org/10.1016/j.chemosphere.2019.03.069.
microbial biomass and plant growth and suppresses Fusarium wilt in tomato. Trop. Yang, X., Liu, J., McGrouther, K., Huang, H., Lu, K., Guo, X., He, L., Lin, X., Che, L.,
Plant Pathol. 45, 73–83. https://doi.org/10.1007/s40858-020-00332-1. Ye, Z., Wang, H., 2016. Effect of biochar on the extractability of heavy metals (Cd,
Soinne, H., Hovi, J., Tammeorg, P., Turtola, E., 2014. Effect of biochar on phosphorus Cu, Pb, and Zn) and enzyme activity in soil. Environ. Sci. Pollut. Res. 23, 974–984.
sorption and clay soil aggregate stability. Geoderma 219–220, 162–167. https://doi. https://doi.org/10.1007/s11356-015-4233-0.
org/10.1016/j.geoderma.2013.12.022. Yao, Y., Gao, B., Inyang, M., Zimmerman, A.R., Cao, X., Pullammanappallil, P., Yang, L.,
Streubel, J.D., Collins, H.P., Tarara, J.M., Cochran, R.L., 2012. Biochar produced from 2011a. Biochar derived from anaerobically digested sugar beet tailings:
anaerobically digested fiber reduces phosphorus in dairy lagoons. J. Environ. Qual. characterization and phosphate removal potential. Bioresour. Technol. 102,
41, 1166–1174. https://doi.org/10.2134/jeq2011.0131. 6273–6278. https://doi.org/10.1016/j.biortech.2011.03.006.
Stumm, W.A., Schindler, P.W., 1987. Aquatic surface chemistry: chemical processes at Yao, Y., Gao, B., Inyang, M., Zimmerman, A.R., Cao, X., Pullammanappallil, P., Yang, L.,
the particle-water interface. In: Stumm, W. (Ed.), The Surface Chemistry of Oxides, 2011b. Removal of phosphate from aqueous solution by biochar derived from
Hydroxides, and Oxide Minerals. John Wiley and Sons, New York, USA. anaerobically digested sugar beet tailings. J. Hazard Mater. 190, 501–507. https://
Takaya, C.A., Fletcher, L.A., Singh, S., Anyikude, K.U., Ross, A.B., 2016. Phosphate and doi.org/10.1016/j.jhazmat.2011.03.083.
ammonium sorption capacity of biochar and hydrochar from different wastes. Yao, Y., Gao, B., Zhang, M., Inyang, M., Zimmerman, A.R., 2012. Effect of biochar
Chemosphere 145, 518–527. https://doi.org/10.1016/J. amendment on sorption and leaching of nitrate, ammonium, and phosphate in a
CHEMOSPHERE.2015.11.052. sandy soil. Chemosphere 89, 1467–1471. https://doi.org/10.1016/J.
Tamburini, F., Pfahler, V., Sperber, C.V., Frossard, E., Bernasconi, S.M., 2014. Oxygen CHEMOSPHERE.2012.06.002.
isotopes for unraveling phosphorus transformations in the soil-plant system: a Yao, Y., Gao, B., Chen, J., Yang, L., 2013a. Engineered biochar reclaiming phosphate
review. Soil Sci. Soc. Am. J. 78, 38–46. https://doi.org/10.2136/ from aqueous solutions: mechanisms and potential application as a slow-release
sssaj2013.05.0186dgs. fertilizer. Environ. Sci. Technol. 47, 8700–8708. https://doi.org/10.1021/
Tan, Z., Zou, J., Zhang, L., Huang, Q., 2018. Morphology, pore size distribution, and es4012977.
nutrient characteristics in biochars under different pyrolysis temperatures and Yao, Y., Gao, B., Chen, J., Zhang, M., Inyang, M., Li, Y., Alva, A., Yang, L., 2013b.
atmospheres. J. Mater. Cycles Waste Manag. 20, 1036–1049. https://doi.org/ Engineered carbon (biochar) prepared by direct pyrolysis of Mg-accumulated tomato
10.1007/s10163-017-0666-5. tissues: characterization and phosphate removal potential. Bioresour. Technol. 138,
Tao, X., Huang, T., Lv, B., 2020. Synthesis of Fe/Mg-biochar nanocomposites for 8–13. https://doi.org/10.1016/j.biortech.2013.03.057.
phosphate removal. Materials 13, 816. https://doi.org/10.3390/ma13040816. Yin, Q., Zhang, B., Wang, R., Zhao, Z., 2017. Biochar as an adsorbent for inorganic
Tao, Y., Han, S., Zhang, Q., Yang, Y., Shi, H., Akindolie, M.S., Jiao, Y., Qu, J., Jiang, Z., nitrogen and phosphorus removal from water: a review. Environ. Sci. Pollut. Res. 24
Han, W., Zhang, Y., 2020. Application of biochar with functional microorganisms for (34), 26297–26309. https://doi.org/10.1007/s11356-017-0338-y.
enhanced atrazine removal and phosphorus utilization. J. Clean. Prod. 257, 120535. Yin, Q., Ren, H., Wang, R., Zhao, Z., 2018. Evaluation of nitrate and phosphate
https://doi.org/10.1016/j.jclepro.2020.120535. adsorption on Al-modified biochar: influence of Al content. Sci. Total Environ.
Tejerina, V.M.R., 2010. Biochar as a Strategy for Sustainable Land Management, Poverty 631–632, 895–903. https://doi.org/10.1016/j.scitotenv.2018.03.091.
Reduction and Climate Change Mitigation/adaptation? MSc Thesis. Faculteit der Yuan, H., Lu, T., Wang, Y., Chen, Y., Lei, T., 2016. Sewage sludge biochar: nutrient
Aard- en Levenswetenschappen Vrije Universiteit, Amsterdam. ECN, p. 71. composition and its effect on the leaching of soil nutrients. Geoderma 267, 17–23.
Thies, J.E., Rillig, M.C., Graber, E.R., 2015. Biochar effects on the abundance, activity https://doi.org/10.1016/j.geoderma.2015.12.020.
and diversity of the soil biota. In: Lehmann, J., Joseph, S. (Eds.), Biochar for Zhai, L., Cai, Z.J., Liu, J., Wang, H., Ren, T., Gai, X., Xi, B., Liu, H., 2015. Short-term
Environmental Management: Science, Technology and Implementation. Routledge, effects of maize residue biochar on phosphorus availability in two soils with different
London, pp. 327–389. phosphorus sorption capacities. Biol. Fertil. Soils 51, 113–122. https://doi.org/
Tomczyk, A., Sokołowska, Z., Boguta, P., 2020. Biochar physicochemical properties: 10.1007/s00374-014-0954-3.
pyrolysis temperature and feedstock kind effects. Rev. Environ. Sci. Biotechnol. 19 Zhang, H., Chen, C., Gray, E.M., Boyd, S.E., Yang, H., Zhang, D., 2016. Roles of biochar in
(1), 191–215. https://doi.org/10.1007/s11157-020-09523-3. improving phosphorus availability in soils: a phosphate adsorbent and a source of
Trazzi, P.A., Leahy, J.J., Hayes, M.H., Kwapinski, W., 2016. Adsorption and desorption of available phosphorus. Geoderma 276, 1–6. https://doi.org/10.1016/j.
phosphate on biochars. J. Environ. Chem. Eng. 4, 37–46. https://doi.org/10.1016/j. geoderma.2016.04.020.
jece.2015.11.005. Zhao, S.X., Ta, N., Wang, X.D., 2017. Effect of temperature on the structural and
Volceanov, E., Georgescu, M., Volceanov, A., Mihalache, F., 2002. Zirconium phosphate physicochemical properties of biochar with apple tree branches as feedstock
binder for periclase refractories. Key Engineering Materials. Trans Tech Publication. material. Energies 10, 1293. https://doi.org/10.3390/en10091293.
Wan, S., Wang, S., Li, Y., Gao, B., 2017. Functionalizing biochar with Mg–Al and Mg–Fe Zheng, Y., Wang, B., Wester, A.E., Chen, J., He, F., Chen, H., Gao, B., 2019. Reclaiming
layered double hydroxides for removal of phosphate from aqueous solutions. J. Ind. phosphorus from secondary treated municipal wastewater with engineered biochar.
Eng. Chem. 47, 246–253. https://doi.org/10.1016/j.jiec.2016.11.039. Chem. Eng. J. 362, 460–468. https://doi.org/10.1016/j.cej.2019.01.036.
Wang, T., Camps-Arbestain, M., Hedley, M., Bishop, P., 2012. Predicting phosphorus Zheng, Q., Yang, L., Song, D., Zhang, S., Wu, H., Li, S., Wang, X., 2020. High adsorption
bioavailability from high-ash biochars. Plant Soil 357, 173–187. https://doi.org/ capacity of Mg–Al-modified biochar for phosphate and its potential for phosphate
10.1007/s11104-012-1131-9. interception in soil. Chemosphere 259, 127469. https://doi.org/10.1016/j.
Wang, T., Camps-Arbestain, M., Hedley, M., 2014. The fate of phosphorus of ash-rich chemosphere.2020.127469.
biochars in a soil-plant system. Plant Soil 375, 61–74. https://doi.org/10.1007/ Zhu, N., Yan, T., Qiao, J., Cao, H., 2016. Adsorption of arsenic, phosphorus and
s11104-013-1938-z. chromium by bismuth impregnated biochar: adsorption mechanism and depleted
Wang, S., Gao, B., Zimmerman, A.R., Li, Y., Ma, L., Harris, W.G., Migliaccio, K.W., 2015. adsorbent utilization. Chemosphere 164, 32–40. https://doi.org/10.1016/j.
Physicochemical and sorptive properties of biochars derived from woody and chemosphere.2016.08.036.
herbaceous biomass. Chemosphere 134, 257–262. https://doi.org/10.1016/J. Zhu, J., Li, M., Whelan, M., 2018. Phosphorus activators contribute to legacy phosphorus
CHEMOSPHERE.2015.04.062. availability in agricultural soils: a review. Sci. Total Environ. 612, 522–537. https://
Wang, Z., Shen, D., Shen, F., Li, T., 2016. Phosphate adsorption on lanthanum loaded doi.org/10.1016/j.scitotenv.2017.08.095.
biochar. Chemosphere 150, 1–7. https://doi.org/10.1016/j. Zhu, D., Chen, Y., Yang, H., Wang, S., Wang, X., Zhang, S., Chen, H., 2020. Synthesis and
chemosphere.2016.02.004. characterization of magnesium oxide nanoparticle-containing biochar composites for
Wang, Y., Ma, Z., Wang, X., Sun, Q., Dong, H., Wang, G., Chen, X., Yin, C., Han, Z., efficient phosphorus removal from aqueous solution. Chemosphere 247, 125847.
Mao, Z., 2019. Effects of biochar on the growth of apple seedlings, soil enzyme https://doi.org/10.1016/j.chemosphere.2020.125847.

13

You might also like